Download CYTOKINE-MEDIATED REGULATION OF BK VIRUS REPLICATION

Document related concepts

Cell encapsulation wikipedia , lookup

Cell cycle wikipedia , lookup

Cell culture wikipedia , lookup

Organ-on-a-chip wikipedia , lookup

Cellular differentiation wikipedia , lookup

Amitosis wikipedia , lookup

List of types of proteins wikipedia , lookup

Interferon wikipedia , lookup

JADE1 wikipedia , lookup

Transcript
CYTOKINE-MEDIATED REGULATION OF BK VIRUS REPLICATION
by
Johanna Renee Abend
A dissertation submitted in partial fulfillment
of the requirements for the degree of
Doctor of Philosophy
(Microbiology and Immunology)
in The University of Michigan
2008
Doctoral Committee:
Professor Michael J. Imperiale, Chair
Professor Dennis Keith Bishop
Associate Professor Kathleen L. Collins
Assistant Professor Scott E. Barolo
Assistant Professor David Miller
ACKNOWLEDGEMENTS
I would like to start by thanking my first mentor, Dr. Stephen Dewhurst at the
University of Rochester, who saw my tiny résumé in his mailbox and gave me a shot in
his lab as an inexperienced sophomore. I would also like to thank Dr. Julie Richards,
who trained me during her graduate career. Working with Steve, Julie, and the rest of the
Dewhurst lab was an amazing experience and gave me a huge head start in research.
Next, I would like to thank Dr. Mike Imperiale for mentoring me throughout my
graduate career. Under his watch, I was challenged scientifically, examined critically,
and always given opportunities to learn and do more. I am grateful for everything.
And how could I ever have made it without my Imperiale labmates? We laughed
together, drank together, thought really hard about science together … Thank you for the
support, thank you for imparting as much knowledge upon me as I could absorb, and
thank you for making me laugh when all my experiments failed in the same hour.
To my family, my favorite and loudest cheering section: thank you for always
believing in me. I cannot imagine what the past five years would have been like without
your constant stream of love and support.
Finally, I would like to thank Adam for keeping me grounded, for listening when
I needed to be heard, and for almost always being the patient one. Thank you for forcing
me to rethink what is most important in life.
ii
TABLE OF CONTENTS
ACKNOWLEDGEMENTS
ii
LIST OF FIGURES
iv
ABSTRACT
vi
CHAPTER I
INTRODUCTION
1
CHAPTER II
INHIBITORY EFFECT OF INTERFERONGAMMA ON BK VIRUS GENE EXPRESSION
AND REPLICATION
41
CHAPTER III
TRANSFORMING GROWTH FACTOR-BETAMEDIATED REGULATION OF BK VIRUS GENE
EXPRESSION
72
CHAPTER IV
PRELIMINARY RESULTS ON THE
CHARACTERIZATION OF INTERFERONGAMMA-MEDIATED REGULATION AND
ARCHETYPE BK VIRUS REPLICATION IN
A TISSUE CULTURE SYSTEM
96
CHAPTER V
DISCUSSION
129
iii
LIST OF FIGURES
Figure
1.1
The genome of BKV
4
1.2
The life cycle of BKV
9
1.3
The structure of archetype NCCR
15
2.1
Dose-dependent IFN-γ inhibition of BKV gene expression
52
2.2
BKV replication kinetics in the presence of IFN-γ
55
2.3
Effect of IFN-γ on viral early region transcript levels
57
2.4
Effect of IFN-γ during infections with different MOIs
59
2.5
Response of various BKV strains to IFN-γ treatment
62
3.1
TGF-β upregulates BKV TU gene expression during infection
81
3.2
BKV early promoter activity in the presence of TGF-β
83
3.3
Alignment of BKV NCCRs
85
3.4
Regions of the TU promoter required for TGF-β-mediated
87
regulation of BKV
4.1
IFN-γ does not affect delivery of BKV DNA to the nucleus
107
4.2
IFN-γ-mediated inhibition of viral gene expression is sustained
108
out to late stages of infection
iv
4.3
Pretreatment with IFN-γ results in greater inhibition of gene
110
expression
4.4
Treatment with HDAC inhibitors restores BKV gene expression
112
and replication in the presence of IFN-γ
4.5
Archetype BKV does not productively infect RPTE cells
119
4.6
Rearranged NCCR can promote archetype BKV DNA replication
121
4.7
Ectopic expression of TAg can facilitate archetype BKV replication
123
5.1
Proposed model for replication of archetype virus and factors that
150
determine the prevalence of rearranged virus in the blood
v
ABSTRACT
BK virus (BKV) is a member of the polyomavirus family that infects nearly the
entire human population at an early age. Following a subclinical primary infection, BKV
is able to establish a persistent infection in the kidney and urinary tract. Reactivation of
BKV occurs in immunocompromised individuals and can lead to severe disease. BKVassociated diseases include polyomavirus nephropathy (PVN), a form of acute interstitial
nephritis that afflicts up to 10% of renal transplant recipients, and hemorrhagic cystitis
(HC), an infection of the bladder characterized by inflammation and hematuria that
affects 10% of bone marrow transplant recipients. PVN and HC are increasing in
prevalence, likely due to the development of more potent immunosuppressive therapies.
To better understand which components of the immune system are important for
regulating BKV infection, we began to examine the effect of cytokines on virus
replication. Interferon-gamma (IFN-γ) has a strong inhibitory effect on BKV
transcription, gene expression, and replication, but does not affect the kinetics of the viral
life cycle or trafficking of the virus to the nucleus. IFN-γ treatment inhibited gene
expression of three different BKV strains similarly, suggesting that regulation by this
cytokine is relevant for all viral strains. We have begun to further examine regulation of
viral transcription by IFN-γ to identify the specific factors involved and investigate
changes in viral chromatin structure. In contrast to IFN-γ, TGF-β has an upregulatory
vi
effect on BKV strain TU early promoter activity. Three other strains examined, however,
were either unaffected or downregulated by TGF-β treatment in a cell type-dependent
manner. The TGF-β response elements were mapped within the promoter of BKV TU.
The viral strain- and cell type-dependent effects of TGF-β demonstrate the complex
nature in which BKV is regulated by cytokine signaling.
Finally, we have begun to investigate the inability of archetype BKV strains to
replicate in tissue culture. The linear structure of the non-coding control region (NCCR)
distinguishes archetype from rearranged strains, which have duplications and deletions of
NCCR sequences. Exchanging the NCCRs of archetype and rearranged genomes restores
archetype DNA replication but prevents rearranged DNA replication. In addition,
archetype DNA replication is observed in the presence of a TAg-expressing plasmid,
suggesting that archetype viruses are limited in replication ability by their lower
production of TAg. Overall, our findings will help us to better understand BKV
persistence in healthy individuals and reactivation in immunocompromised patients.
vii
CHAPTER I
INTRODUCTION
Viruses have long been studied for their ability to evade the host immune
response and carry out their life cycles in the context of otherwise healthy individuals.
For many viruses, this is primarily mediated by the production of viral gene products that
mimic or block the mounting immune response. Encoding such proteins in the viral
genome, however, requires genetic space and therefore small viruses are limited in their
capacity to make these factors. BK virus (BKV) has a genome of only 5.2 kb and
encodes only seven proteins, yet is able to establish a persistent infection in healthy
individuals. This chapter will provide detailed information on the molecular biology,
immunology, and clinical aspects of BKV to provide the necessary background for the
discussion of viral regulation by the host immune response and establishment of
persistent infection in the following chapters.
1
Genetics and Life Cycle of BK Virus
BKV is a member of the family Polyomaviridae, a group of viruses characterized
by similar genome structures, small nonenveloped virions, and the ability to induce
tumors in cells that do not support productive infections (Imperiale and Major, 2007).
There are two known human polyomaviruses, BKV and JC virus (JCV), however recent
reports of viral sequences isolated from respiratory samples and skin cancer cells suggest
the existence of three additional human polyomaviruses, WU, KI, and Merkel cell
polyomavirus (Allander et al., 2007; Feng et al., 2008; Gaynor et al., 2007). The
polyomavirus family also includes the well-studied simian virus 40 (SV40) and mouse
polyomavirus (Py); many details of BKV molecular biology are derived from what is
known for these two viruses. BKV is highly homologous to JCV and SV40, sharing 78
to 90% homology at the amino acid level over the major viral proteins, and 75% and 69%
DNA sequence homology throughout the genome, respectively (Cubitt, 2006; Imperiale,
2001). The polyomaviruses, however, are highly species specific and can only replicate
efficiently in their natural host cells. In addition, there is evidence suggesting the
coevolution of these viruses with their hosts (Shadan and Villarreal, 1993). Thus, SV40
and Py can only provide limited information about the immune response to BKV or the
dynamics of infection, transmission, and persistence in humans.
BKV has a 40 to 45 nm icosahedral virion with a T = 7 lattice symmetry (Li et al.,
2003). The virion is composed of 360 copies of VP1, arranged in 72 capsomeres, each of
which contains five copies of VP1 and one copy of either VP2 or VP3 (Imperiale and
Major, 2007). The genome is a circular, double-stranded DNA approximately 5 kb in
length, and is associated with cellular histones H2A, H2B, H3 and H4 both in the nucleus
2
of infected cells and in the virion (Meneguzzi et al., 1978). This association with cellular
histones leaves the virus vulnerable to regulation by host chromatin remodeling agents.
The genome of BKV has three major regions: the non-coding control region, the
early coding region, and the late coding region (Figure 1.1). The non-coding control
region (NCCR) contains a bidirectional promoter that drives the production of early
region transcripts from one genomic strand and late region transcripts from the opposite
strand. In addition, the NCCR contains the origin of replication, three TAg binding
motifs, and many cellular transcription factor binding sites. The NCCR is the region of
greatest variability between different isolates and thus distinguishes one “strain” of BKV
from another (NCCR variants are henceforth termed strains in accordance with the
literature). The NCCR is also the region of least conservation among members of the
polyomavirus family. The early coding region contains the genes for large tumor antigen
(TAg) and small tumor antigen (tAg), the transcripts of which are produced by alternative
splicing of a common mRNA precursor. There is now evidence of a third early protein in
BKV called truncated T antigen (truncTAg; D. Das, A. Joseph, J. Abend, D. CampbellCecen, and M. Imperiale, in preparation), that has a transcript structure similar to that of
SV40 17kT and the T’ proteins of JCV (Trowbridge and Frisque, 1995; Zerrahn et al.,
1993). The early region is highly conserved between the different BKV strains. The late
coding region contains the genes for viral capsid proteins VP1, VP2, and VP3, and the
agnoprotein, the transcripts of which are also all produced by alternative splicing of a
common mRNA precursor. VP2 and VP3 are translated in the same reading frame using
different start codons, such that VP3 is an N-terminal truncated form of VP2. VP1 is
translated from the same transcript but using a start codon in a different reading frame
3
Figure 1.1. The genome of BKV. This schematic shows the circular, double-stranded
DNA genome of BKV. The three major regions are shown as follows: the non-coding
control region (NCCR, bottom section), the early coding region (left hemisphere), and the
late coding region (right hemisphere). Solid arrows represent transcripts that encode the
viral proteins; dashed lines represent the differential splicing of the early coding region.
Figure courtesy of Mengxi Jiang.
4
than that of VP2 and VP3. The 5’ end of the transcript has another open reading frame
from which agnoprotein is translated. While the late proteins are highly conserved, there
is a short region of variability (amino acids 61 to 83) within VP1 that is used to assign
each strain of BKV to a particular subtype (I, II, III, IV; Jin, 1993; Jin et al., 1993).
Subtype I viruses predominate in the human population, followed by subtype IV, while
subtypes II and III are rare (Cubitt, 2006; Nukuzuma et al., 2006). The functional
relevance of BKV subtypes is not yet fully understood.
The pathway for BKV entry and trafficking to the cell nucleus is not yet fully
understood, but is somewhat distinct from that of JCV, SV40, and Py. The receptor for
BKV is either of the gangliosides GD1b and GT1b (Low et al., 2006) or an N-linked
glycoprotein with an α(2,3)-linked sialic acid (Dugan et al., 2005); both reports note the
importance of sialic acids in BKV receptor binding. Upon engaging the receptor, the
virus is internalized in a caveolae-dependent manner, similar to SV40 but distinct from
JCV, which enters by clathrin-dependent endocytosis, and Py, which enters via a unique
endocytic pathway (Anderson et al., 1996; Atwood, 2001; Eash et al., 2004; Pho et al.,
2000). The subsequent steps of BKV trafficking to the nucleus are currently under
investigation, but the pathway includes passage of the virus through the endoplasmic
reticulum and involves both microtubules and actin filaments (Eash and Atwood, 2005;
Low et al., 2006; Moriyama and Sorokin, 2008). It is still unclear whether the partially
disassembled virion or only the viral minichromosome is transported into the nucleus
where replication occurs.
The first event to follow delivery of the viral genome to the nucleus is
transcription of the early region and subsequent production of TAg and tAg. The
5
processes of transcription, translation, and DNA replication for BKV require the use of
host cell enzymatic machinery, as the viral genome does not encode the proteins to
accomplish these tasks. TAg is arguably the most important of the viral proteins as it
assists and regulates the host cell machinery in these processes, performing many
functions that are critical to the viral life cycle (based on findings for SV40; Cole and
Conzen, 2001; Kierstad and Pipas, 1996; Kim et al., 2001; Moens and Rekvig, 2001).
Once translated in the cytoplasm, TAg translocates back to the nucleus by means of a
highly conserved nuclear localization signal at the N-terminus. TAg then interacts with
the pRB family of proteins (pRB, p107, and p130), which releases the E2F transcription
factor and drives the cell into S phase (Harris et al., 1996; Harris et al., 1998). There are
three known binding sites for TAg at the origin of replication which mediate several
functions: 1) inhibition of early gene transcription by blocking RNA polymerase binding,
2) initiation of viral genome replication by recruiting cellular proteins and acting as a
helicase to facilitate the unwinding of the genome (Stahl et al., 1986; Stillman et al.,
1985), and 3) activation and promotion of late gene transcription by binding to the late
promoter (Deyerle et al., 1989; Salzman et al., 1986). TAg also binds and inactivates the
tumor suppressor p53 to promote cell survival during infection (Harris et al., 1996;
Shivakumar and Das, 1996). Overall, TAg establishes a cellular environment that is
supportive of viral replication. Due to its importance to the viral life cycle, TAg
expression may also be a key point of regulation by the host immune system.
Small T antigen is somewhat viewed as an auxiliary TAg and is distributed
throughout the cytoplasm and nucleus (Cole and Conzen, 2001). It has been shown to
enhance DNA replication (Cicala et al., 1994), mediate cell cycle progression (Howe et
6
al., 1998; Whalen et al., 1999), and aid in transformation of cells (Manfredi and Prives,
1994). A major function of tAg involves interaction with and inactivation of protein
phosphatase 2A, which leads to cell proliferation and transformation; this is a unique
function of tAg not shared by TAg (Skoczylas et al., 2004; Yang et al., 1991). It appears,
however, that tAg is dispensable in the context of lytic infection, based on the ability of
BKV strain MM to mediate a productive infection in the absence of detectable tAg
expression (Seif et al., 1979).
The production of early proteins and the establishment of a suitable host cell
environment promote replication of the viral genome. BKV DNA replication begins at
the origin of replication, proceeds bidirectionally using host DNA polymerase α-primase
to prime the extension by DNA polymerase δ, and ends at the opposite side of the
genome, where DNA ligase I replaces missing nucleotides and topoisomerase II mediates
the separation of the daughter molecules (based on SV40 replication; reviewed in Kim et
al., 2001). Late gene transcription occurs concurrently and the capsid proteins (VP1,
VP2, VP3) are translated and translocated back to the nucleus, the site of viral assembly.
Lastly, agnoprotein is expressed at very late stages of lytic infection. BKV agnoprotein is
localized to the cytoplasm and perinuclear region, and interacts with cellular proteins of
various molecular weights (Rinaldo et al., 1998). Although the function is not welldefined, it has been suggested that agnoprotein is important for virus assembly,
maturation, and/or release (Rinaldo et al., 1998).
Accumulation of late proteins and replicated genomes in the nucleus leads to
virion assembly and release of progeny viral particles from the infected cell. The details
of assembly and release have not been determined. This is the completion of the viral life
7
cycle in a permissive cell (Figure 1.2), resulting in production of mature, infectious
particles and concluding with lysis of the infected cell or an undetermined mechanism of
viral egress to release the progeny virions. In addition, BKV can infect cells that are
nonpermissive for productive infection, most notably rodent cells. Nonpermissive
infection results in early gene expression (TAg, tAg) but a failure to replicate viral DNA,
most likely due to the inability of TAg to interact with and recruit cell replication
machinery of another species. The accumulation of TAg and tAg without progression to
the late stages of infection results in oncogenic transformation.
BK Virus Persistence and Associated Disease
BKV was first discovered in 1971 in the urine of a renal transplant patient with
ureteral stenosis (Gardner et al., 1971). In the same year, JCV was isolated from human
brain tissue from a patient with progressive multifocal leukoencephalopathy (PML)
(Padgett et al., 1971). From the beginning of their acknowledged existence, the human
polyomaviruses were linked to significant disease.
The route of transmission for BKV is not well-defined. It is assumed that primary
infection occurs either by the oral or respiratory route because of the widespread nature
of BKV infection and the timing of seroconversion: 50% of children are seropositive by
the age of 3 and 91% are seropositive by the age of 9 (Knowles et al., 2003). The
primary infection is typically subclinical, but has been associated with upper respiratory
disease in children (Goudsmit et al., 1982). Approaching adulthood, BKV
seroprevalence declines slightly to approximately 81% for the general population
(Knowles et al., 2003). Following primary infection, BKV disseminates throughout the
8
Figure 1.2. The life cycle of BKV. Numbers indicate the steps of BKV infection and
replication: 1) binding to receptor GD1b, GT1b, or sialic acid; 2) caveolae-dependent
endocytosis; 3) trafficking, which involves microtubules and actin filaments; 4) passage
through the endoplasmic reticulum; 5) delivery of genome to the nucleus; 6) early gene
transcription and translation; 7) TAg returns to the nucleus to facilitate viral DNA
replication and late gene transcription; 8) late gene transcription and translation; 9) capsid
proteins return to the nucleus for assembly of progeny virions; 10) release of viral
progeny by egress or cell lysis.
9
body; viral sequences have been reported in a variety of different cell types, including
peripheral blood mononuclear cells (Chatterjee et al., 2000; Doerries et al., 1994), liver
(Knepper and diMayorca, 1987), bone tissue (De Mattei et al., 1994), and brain (Elsner
and Doerries, 1992). Finally, BKV reaches the cells of the kidney and urinary tract,
specifically proximal tubule epithelial cells and the urothelium (Chesters et al., 1983;
Heritage et al., 1981), where it establishes a persistent, lifelong subclinical infection in
healthy individuals. It is reported that 50% of healthy native kidneys and 40% of ureters
harbor BKV, with the highest levels (400 genome copies per 100,000 cells) in the renal
medulla (Monini et al., 1995; Nickeleit et al., 2003). Approximately 5% of healthy
individuals actively excrete BKV in urine (viruria) at a rate of greater than 106 genome
copies per ml (Zhong et al., 2007). In addition, there is a high tendency of pregnant
women to be viruric with no clinical outcome (Coleman et al., 1980; Markowitz et al.,
1991). In healthy individuals, BKV has developed an effective means to coexist with the
host, maintaining a subclinical infection that sporadically leads to viral excretion in the
urine and potential transmission to new hosts.
Certain conditions of the host, specifically immunosuppression, can allow BKV to
‘reactivate’ from the persistent subclinical state to an active lytic infection, resulting in
serious disease. This is a feature common to polyomaviruses, as JCV, Py, and SV40 all
undergo reactivation in immunocompromised hosts (Imperiale and Major, 2007). Two
major patient groups are highly susceptible to BKV reactivation: bone marrow transplant
(BMT) recipients and kidney transplant (KT) recipients. In BMT patients, BKV lytic
infection occurs in the epithelial cells of the bladder or ureter, resulting in hemorrhagic
cystitis (HC) characterized by viruria and hematuria. HC is reported in 10 to 68% of
10
BMT patients, with 25% suffering from severe HC with heavy bleeding and intense pain
(Arthur et al., 1986; Azzi et al., 1994; Bedi et al., 1995; Cotterill et al., 1992). BKVassociated HC occurs more than 10 days post-transplant (late-onset) and requires medical
attention but is not usually life-threatening (Egli et al., 2007). It has been suggested that
HC is an immune reconstitution syndrome, in which the reactivation of BKV, and thus
the presence of large amounts of viral antigens, sets off an excessive inflammatory
response that causes damage to the urothelial cells (Hirsch, 2005).
In KT recipients, reactivation of BKV is first evident by viral shedding in the
urine, which occurs in 20 to 60% of patients. About half of these patients then develop
high levels of viremia (virus dissemination in the bloodstream; Brennan et al., 2005;
Hirsch, 2005; Hirsch et al., 2002). In these patients, BKV that established a subclinical
infection in kidney epithelial cells has reactivated to a productive lytic infection in which
viral progeny have spread to the surrounding cells of the medulla, renal cortex, and
proximal tubules. This condition is referred to as polyomavirus nephropathy (PVN) and
the end stage involves necrosis of the tubules and denudation of the basement membrane
(Nickeleit et al., 2003). PVN is diagnosed primarily by examination of kidney biopsies
for characteristics of virus-infected cells (enlarged nuclei, cytopathic effects), followed
by immunohistochemistry to detect TAg expression (Drachenberg et al., 2005; Vats et al.,
2006). In addition, PCR quantitation of viral genomes in biopsy samples and urine
cytology to identify decoy cells (BKV-infected cells with characteristic morphology) are
often employed (Singh et al., 2006). If not controlled, the lytic infection will cause
failure and destruction of the graft. Approximately 1 to 10% of KT patients develop
PVN, primarily during the first year post-transplantation, which coincides with the time
11
of highest immunosuppressive treatments. Once diagnosed with PVN, 50 to 90% of
patients progress to graft failure (Egli et al., 2007). The incidence of PVN is increasing,
most likely due to the combined effects of more potent immunosuppressive therapies,
better methods of diagnosis, and the increasing number of KT performed every year.
The treatment options for PVN are quite limited. The most effective treatment to
date is reduction of the immunosuppressive regimen, which allows the patient to mount
an immune response and fight the infection. There are currently no proven antiviral
treatments for BKV, although some limited success has been seen with cidofovir and
leflunomide (Bernhoff et al., 2008; Vats et al., 2006). A third approach is nephrectomy
to remove the transplanted kidney. This has been shown to cause a rapid decrease in
viral load, with a viral half-life as fast as one to two hours and greater than 99% turnover
each day, compared to a viral half-life of six hours to 17 days resulting from a decrease in
immunosuppression (Funk et al., 2006). These findings suggest that the graft is the
source of replicating BKV, not the kidney epithelial cells or urothelium of the recipient.
There is no single risk factor shared by all patients that develop PVN; it is thought
that BKV reactivation is prompted by a combination of factors from the virus, patient and
graft (Comoli et al., 2006; Egli et al., 2007; Hirsch et al., 2006). Specific individual risk
factors include older age, male gender, seropositivity of the donor, seronegativity of the
recipient, specific immunosuppressive drugs, HLA mismatches, and acute rejection
episodes prior to development of PVN (Andrews et al., 1988; Bohl et al., 2005; Comoli et
al., 2006; Egli et al., 2007; Mengel et al., 2003). It is clear that an immunosuppressed
state alone is not enough to cause PVN; not all KT patients suffer from PVN and there
are only sporadic references to BKV reactivation and associated nephropathy in other
12
solid organ transplant patients or immunodeficiency syndromes, such as AIDS (Hirsch,
2005; Pavlakis et al., 2006).
Therefore, a primary topic addressed by investigators currently studying BKV
reactivation is the identification of component(s) of the immune system that are critical
for controlling BKV infections. Further knowledge about the immune response to and
control of BKV in healthy individuals may allow better treatment options for patients
with PVN and HC. Several clinical observations have been made about the immune
response during BKV reactivation. It is reported that 77% of patients are seropositive for
BKV before KT (Hirsch et al., 2002) and that patients with detectable anti-BKV
antibodies still progress to PVN (Comoli et al., 2004). In fact, it seems that antibodies
are more indicative of the viral load in a patient rather than protective against BKV
infection, as high levels of antibodies in PVN patients correlate with high levels of
viremia and low CD8+ T cell responses (Chen et al., 2006). Thus, anti-BKV antibodies
do not appear to play a major role in controlling infection and reactivation. Furthermore,
the cell-mediated cytotoxic immune response fails to eliminate all BKV-infected cells,
allowing the establishment of a persistent infection in healthy individuals. Cytokineproducing T cells now show the most promise as key regulators of BKV reactivation: low
levels of BKV-specific IFN-γ-producing T cells correlate with progression to PVN, while
reconstitution of these cells correlates with resolution of PVN (Binggeli et al., 2007;
Chen et al., 2006; Prosser et al., 2008). Therefore, the fluid-borne effectors produced by
T cells, such as cytokines, may be a critical means to control BKV replication and
reactivation.
13
BKV NCCR and Transcriptional Regulation
BKV strains can be divided into two major types, archetype and rearranged, based
on the structure of the NCCR. The archetype NCCR has a simple structure composed of
five blocks of nucleotide sequences, designated O, P, Q, R, and S (Figure 1.3; Markowitz
and Dynan, 1988; Rubinstein et al., 1987; Sundsfjord et al., 1994). Rearranged strains
have major structural changes in the NCCR, relative to the archetype, involving partial or
whole duplications, deletions, or mutations of these blocks. The O, P, Q, R, and S blocks
were defined as sequences that appear to move together when rearrangements occur. The
O block is 142 bp long and highly conserved among all strains of BKV, likely because it
contains the origin of replication, the TATA box, and the three TAg binding sites that are
required to initiate replication. The P block is 68 bp long and is present at least once in
every strain isolated thus far, though it is frequently duplicated in part or in whole
(Moens and Van Ghelue, 2005). This region contains a cAMP response element (CRE)
and proven binding sites for the transcription factors Sp1, AP-1, and NF-1. The Q block
is 39 bp long and contains proven binding sites for Sp1, NF-1, and NF-κB. The R block
is 63 bp long and has only been shown to contain an NF-1 binding site. The S block,
leading up to the start codon of agnoprotein, is 63 bp long and contains two proven NF-1
sites, an estrogen response element, and a glucocorticoid and/or progesterone response
element. The Q and R blocks are frequently deleted from rearranged strains, while the S
block is rarely altered (Cubitt, 2006; Johnsen et al., 1995; Moens and Rekvig, 2001;
Moens and Van Ghelue, 2005).
Rearrangements of the NCCR can add or remove transcription factor binding sites
either by the addition or deletion of blocks that contain these sites, or by splicing
14
Figure 1.3. The structure of archetype NCCR. A schematic representation of the
structure of archetype BKV NCCR. Blocks of transcription factor binding sites are
labeled by letter and length. The locations of binding sites discussed in the text are
shown. Open green triangles represent TAg binding sites. Open red circles represent
NF-1 binding sites. Arrows indicate the bidirectional nature of the promoter, initiating
early gene transcription to the left and late gene transcription to the right. Below is a
schematic representation of the structure of a rearranged NCCR, BKV TU, which is used
in most of our experiments. CRE, cAMP response element; HRE, hormone response
element.
15
sequences together, creating new binding sites at the junctions of blocks (Markowitz and
Dynan, 1988; Markowitz et al., 1990). Furthermore, BKV promoters may be regulated
differently depending on the cell type infected. It has been shown by DNase footprint
analysis that different protein complexes bind to the NCCR during incubation with HeLa
versus 293 cell extracts (Grinnell et al., 1988). The observed binding patterns correlate
with differences in promoter activity in these two cell lines, suggesting that replication
capacity may also differ in a cell type-dependent manner.
From the beginning, it was clear that there is a functional difference between
rearranged and archetype NCCRs. Archetype viruses are more efficient at transforming
rodent cells but extremely inefficient at propagation in tissue culture, while rearranged
BKV strains are highly adapted to growth in tissue culture but unable to efficiently
transform cells (Watanabe and Yoshiike, 1982; Watanabe and Yoshiike, 1986). Upon
examination of the individual blocks, it was observed that an NCCR with a single P block
was enough to initiate transcription, but that repetitions of the P block resulted in greater
promoter activity (studies in HeLa cells; Chakraborty and Das, 1989; Deyerle and
Subramani, 1988). Insertion of P block fragments of a rearranged NCCR into an
archetype NCCR resulted in enhanced transcriptional activity, indicating that it is not the
presence of repressor elements in archetype strains, but instead the duplication of
activator elements in rearranged strains that results in the elevated capacity to replicate
(Markowitz et al., 1990). By comparing the growth phenotypes of different BKV strains
to their NCCR structures, it was noted specifically that the duplication of P blocks favors
growth in HeLa cells (Markowitz and Dynan, 1988). It was also shown that, in the
presence of HeLa cell extracts and immunoaffinity purified TAg, a plasmid containing an
16
NCCR with a single P block replicated as well as one with three adjacent P blocks, and
that mutation of the NF-1 and Sp1 binding sites within the P block had no effect on
replication (Del Vecchio et al., 1989). These results suggest that, in the presence of
abundant TAg, the transcription and replication elements of the NCCR are separable;
however, the functions in vivo are not, since replication requires efficient TAg
expression, implying that the greater the activity of the early promoter, which drives the
expression of TAg, the greater the replication capacity of the virus.
It has long been thought that NCCR structure would play a role in BKV
pathogenesis by giving rearranged viruses selective replication advantage over the
archetype strains. In many independent experiments, an archetype strain from a clinical
isolate showed essentially no replication in tissue culture over the course of many weeks.
If viral replication was detected, the NCCR structure of the progeny virus was
rearranged, indicating that either a rearrangement event had occurred or the clinical
sample used for infection contained an initially undetectable level of rearranged virus that
eventually outgrew the archetype (Rinaldo et al., 2005; Rubinstein et al., 1991;
Sundsfjord et al., 1994; Sundsfjord et al., 1990). These studies have been performed in
human umbilical vein endothelial (HUVEC-C), monkey kidney epithelial (Vero), and
human embryonic kidney (HEK) cell lines. A similar phenomenon is seen when
archetype genomic DNA is transfected into cells: after extended incubation, the progeny
virus, if produced, has NCCR rearrangements (Rinaldo et al., 2005; Rubinstein et al.,
1991). This result directly demonstrates the evolution of a rearranged strain of BKV
from an archetype genome.
17
Despite the evidence that rearranged strains are better at replicating in tissue
culture, the vast majority of strains isolated from both healthy individuals and
immunosuppressed patients have archetype NCCRs (Gosert et al., 2008; Markowitz et
al., 1991; Negrini et al., 1991; Sharma et al., 2007; Sugimoto et al., 1989; Sundsfjord et
al., 1999; Takasaka et al., 2004; ter Schegget et al., 1985). It was recently reported,
however, that rearranged strains are found more frequently in the blood of patients with
PVN than in the urine, and that these strains attain 20-fold higher viral loads than
archetype strains in the blood (Gosert et al., 2008). In addition, kidney biopsies from
PVN patients with rearranged strains had more evidence of inflammation than those with
archetype strains. This is the first report to link rearranged strains directly to more
extensive viral replication and subsequent disease in patients. Still, the authors could not
identify any commonalities in the rearrangements that would implicate a specific NCCR
element in pathogenesis. The authors hypothesized that archetype strains are more
successful in immunocompetent hosts because of their ability to remain relatively
undetected by the immune system, by means of slower growth and lower levels of TAg
(Gosert et al., 2008). While the clinical relevance of NCCR rearrangements is not yet
clear, many researchers are trying to show the association between transcription factor
binding sites or NCCR structures with progression to disease. For example, mutations in
the Sp1 binding site at the junction of the Q and R blocks may be found more frequently
in BMT patients with HC than those without, but the relevance and effect of the mutation
on viral replication is not known (Priftakis et al., 2001).
18
Interferon-Gamma and Polyomaviruses
Interferons are a class of cytokines widely known for their antiviral effects, and in
fact were first discovered and named for their ability to “interfere” with viral infection
(Isaacs and Lindenmann, 1957; Wheelock, 1965). There are three classes of interferons,
type I, type II, and type III, related by function and signaling pathways (reviewed in
Biron and Sen, 2001; Pestka et al., 2004). Type I interferons are a constantly growing
class of molecules, which currently include IFN-α, IFN-β, IFN-ε, IFN-δ, IFN-κ, IFNτ, IFN-ω, and the interferon-like cytokine limitin. Type III interferons, or IFN-λ, were
only recently discovered and have two members, IL-28 and IL-29 (Ank et al., 2006).
IFN-γ is the only member of the type II class, distinct from type I in that it is produced
primarily by immune cells, mainly natural killer (NK) cells, T cells, antigen presenting
cells (APCs), and B cells. Type I interferons play a major role in innate immunity against
viruses, as they are produced by almost all cell types and are quickly induced by viral
components, such as double-stranded RNA. In contrast, IFN-γ can be produced as an
innate immune mediator by NK cells, or as an adaptive immune mediator when produced
by T cells, B cells, and APCs. Viral antigens must be presented to T cells in the proper
MHC context to activate IFN-γ expression. IFN-γ is primarily associated with the T
helper 1 (Th1) phenotype: production is enhanced by IL-2, IL-1, estrogen, IL-18, IL-12,
and IFN-γ itself, but inhibited by glucocorticoids, TGF-β and cytokines associated with
the Th2 phenotype (Biron and Sen, 2001).
IFN-γ is a 21 kDa cytokine expressed as a homodimer. The receptor is present on
the surface of almost all nucleated cells and is composed of two heterodimers of IFNGR1
19
and IFNGR2 molecules, distinct from the IFN-α/β receptor. Janus kinase 1 (JAK1) is
bound to the cytoplasmic tail of IFNGR1 while JAK2 is bound to IFNGR2, and upon
IFN-γ binding, these two kinases are phosphorylated. JAK1 and JAK2 then
phosphorylate the cytoplasmic domain of IFNGR1, which allows binding of signal
transducer and activator of transcription 1 (STAT1) to the receptor. STAT1 molecules
are then phosphorylated which results in their dimerization; homodimers of STAT1,
known as the gamma-activated factor (GAF), translocate to the nucleus and bind gammaactivated sequences (GAS) to promote transcription of IFN-γ responsive genes.
Crosstalk between the IFN-α/β and the IFN-γ signaling cascades can result from IFNα/β-stimulated activation of GAF, although primarily the type I interferons signal
through the complex of STAT1, STAT2, and interferon regulatory factor 9 (IRF-9),
which binds the interferon-stimulated response element (ISRE) in promoters.
Furthermore, in certain situations GAF can complex with IRF-9 and mediate signaling
through ISREs (reviewed in Pestka et al., 2004; van Boxel-Dezaire and Stark, 2007).
The effects of IFN-γ signaling on a cell are far-reaching and diverse (Biron and
Sen, 2007; Sen, 2001; van Boxel-Dezaire and Stark, 2007). First, IFN-γ has an important
immunomodulatory role which it mediates by activating monocytes to produce nitric
oxide synthase 2 (iNOS, important for eliminating bacteria), inducing expression of
chemokines to recruit and activate immune cells, activating NK cell cytotoxicity,
increasing MHC class I and II expression, promoting differentiation of CD4+ T cells, and
promoting immunoglobulin class switching. In nonimmune cells, interferon signaling is
mostly directed at establishing an antiviral state. Among the genes activated are protein
kinase R (PKR), which phosphorylates eukaryotic initiation factor 2 (eIF-2α) to inhibit
20
translation; 2’-5’ oligoadenylate synthetase, which activates RNaseL to nonspecifically
degrade mRNA; adenosine deaminase, which introduces adenosine to inosine changes in
transcripts to affect protein expression and function; and IRFs, which potentiate the
signaling cascade by mediating interferon synthesis and interferon responsive gene
regulation. The overall effect on the cell is anti-proliferative and pro-apoptotic.
Transcription factors other than STAT1 known to be activated during IFN-γ signaling
include STAT3, AP-1, USF-1, NF-κB, IRF-1, IRF-8, ATF-2, GATA-1, CREB, and PU.1,
as well as C-EBP-β and CIITA, which are newly synthesized in response to signaling and
mediate the second wave of the cascade (van Boxel-Dezaire and Stark, 2007). The
crosstalk between the type I and type II interferon pathways creates difficulty in defining
genes as being specifically regulated by IFN-α/β or IFN-γ. Several studies have
examined the differential effects of interferon signaling by microarray in various cell
types, including human lung carcinoma (Sanda et al., 2006; Tan et al., 2005) and human
fibrosarcoma (Der et al., 1998) cell lines.
Many viruses have developed methods to overcome the interferon response by
blocking interferon synthesis or interferon-mediated signaling and effectors, or even by
expressing interferon receptor decoys (Biron and Sen, 2007; Sen, 2001). Despite the
small genome size, and thus the limited ability to encode specific inhibitor and decoy
proteins, there are several reports that polyomaviruses can resist the antiviral effects of
interferons. SV40 TAg can inhibit PKR-mediated deactivation of eIF-2α, most likely by
stimulating a cellular phosphatase to dephosphorylate eIF-2α (Swaminathan et al., 1996).
Py has been shown to specifically inhibit IFN-β-stimulated gene expression by blocking
21
the activation of the STAT1/STAT2/IRF-9 complex through the interaction of TAg with
JAK1 (Weihua et al., 1998).
The regulation of human polyomaviruses by interferons is currently being
investigated. With the increasing popularity of microarray studies, two groups have
separately examined changes induced by JCV infection of cultured cells. The first study
analyzed the effect of infection with JCV/SV40 chimeric viruses, containing JCV coding
regions and JCV/SV40 hybrid NCCRs, on human fetal astrocyte cultures. During
infection, there was a strong upregulation in cell cycle genes, as anticipated, but
surprisingly no effect on interferon-responsive genes and very little effect on genes
relating to the immune response in general (Radhakrishnan et al., 2003). In contrast, the
second study reported that transfection of JCV genomic DNA into primary human fetal
glial cells stimulated a variety of immune response genes, in particular interferonstimulated genes (Verma et al., 2006). The upregulation of several factors was confirmed
during JCV infection and, in a follow-up study, the authors reported that the induction of
the interferon response required active JCV replication, not just the presence of viral
products or DNA (Co et al., 2007). In addition, they noted an inhibition of JCV
replication by IFN-α/β treatment at late time points during infection (Co et al., 2007).
There has only been one microarray study published analyzing the effect of BKV
infection on immortalized human umbilical vein endothelial cells. As expected, the
authors report the upregulation of cell cycle-related genes at various times after infection;
however, they saw no induction of interferon transcripts, the upregulation of only two
interferon-inducible genes, and the downregulation of several genes involved in antiviral
defense (Grinde et al., 2007). These results indicate that BKV may have the ability to
22
infect cells in such a way as to avoid detection by the innate intracellular immune
response.
Transforming Growth Factor-Beta
Members of the TGF-β family of cytokines are secreted proteins that share
sequence and structural features and have similar signaling cascades within the cell. This
family includes the bone morphogenic proteins (BMPs), which are important during
development; activins, which regulate growth and differentiation; and the TGF-β
isoforms, TGF-β1, TGF-β2, and TGF-β3, which have similar functions in the adaptive
immune response. The TGF-β family members are expressed in most cell types,
although the effect on TGF-β signaling can be cell type-dependent. This section will
focus specifically on the details of TGF-β1 signaling and regulation.
The precursor of TGF-β is a prepro-peptide that contains a signal peptide for
secretion, an immature propeptide called the latency associated protein (LAP) that is
removed by proteolytic cleavage in the Golgi (Dubois et al., 1995), and the mature TGFβ protein, a 25 kDa dimer (Feng and Derynck, 2005; Li et al., 2006; Rahimi and Leof,
2007). TGF-β can be secreted in two different forms: the small latent complex (SLC), a
homodimer of mature TGF-β associated with a homodimer of LAP, or the large latent
complex (LLC), the SLC in association with latent-TGF-β binding protein (Annes et al.,
2003). The liberation of mature TGF-β from the latency proteins in vivo is not fully
understood, but can be mediated by thrombospondin-1, matrix metalloproteinases,
integrins and reactive oxygen species, among others (Li et al., 2006; Yang et al., 2007).
The mature TGF-β binds to a cell surface receptor that is composed of two type I
23
receptors (TβRI) and two type II receptors (TβRII). Upon ligand binding, the receptor
complex autophosphorylates: first TβRII phosphorylates TβRI and then TβRI itself
undergoes autophosphorylation. Finally, TβRI phosphorylates Smad proteins in the
cytoplasm (Derynck and Zhang, 2003; Shi and Massague, 2003).
There are eight Smad proteins expressed in mammalian cells, falling into three
classes: the receptor Smads (R-Smad, Smad1, 2, 3, 5, 8), the common Smad (Co-Smad,
Smad4), and the inhibitory Smads (I-Smad, Smad6, 7). R-Smads (Smad2 and 3 in the
case of TGF-β1) are phosphorylated by TβRI, form a complex of two R-Smads and one
Co-Smad, and then translocate to the nucleus to act as transcription factors. This
Smad2/3/4 complex binds to (5’ CAGAC 3’) sequences in promoters, also known as the
Smad binding element (SBE; Shi et al., 1998; Zawel et al., 1998). Smad proteins by
themselves have only weak DNA binding activity unless multiple SBEs are present (Shi
et al., 1998; Zawel et al., 1998). More commonly, the complex interacts with other
transcription factors that have more specific or higher affinity DNA binding abilities to
regulate gene expression. These factors must have the ability to interact with Smad
proteins and have a DNA binding site nearby to the SBE (Derynck and Zhang, 2003; Shi
and Massague, 2003). Co-Smad interactions help to stabilize the R-Smad interactions
with other transcription factors. I-Smads are negative regulators of Smad signaling:
Smad6 competes with Smad4 for binding to R-Smads and forms inactive complexes
(Hata et al., 1998), while Smad7 competes with R-Smads for binding to the activated
receptors (Hayashi et al., 1997; Nakao et al., 1997). Smad7 is induced by IFN-γ
signaling through the JAK/STAT pathway and by NF-κB signaling in response to
inflammatory cytokines (Bitzer et al., 2000; Giannopoulou et al., 2006; Ulloa et al.,
24
1999). There is an extensive list of transcriptional coactivators and corepressors that
interact with Smad complexes and mediate TGF-β-responsive gene regulation, including
AP-1, FAST proteins, NF-κB, p300/CBP, P/CAF, Sp1, and ZEB proteins (Brown et al.,
2007; Feng and Derynck, 2005).
The effects of TGF-β family members vary depending on the cell type, but overall
these cytokines can regulate proliferation, differentiation, migration, and cell survival
during development, carcinogenesis, fibrosis, wound healing and immune responses
(Blobe et al., 2000; Li et al., 2006). In immune cells, TGF-β-mediated signaling
generally results in inhibition of growth and differentiation, resulting in tolerance and
immunosuppressive effects (Li et al., 2006). Similarly, TGF-β signaling in epithelial
cells usually results in growth arrest and apoptosis; in fibroblasts, however, TGF-β
promotes proliferation and activation (Rahimi and Leof, 2007). Several reports have
shown that various immunosuppressive therapies commonly used in both solid organ
transplant and BMT patients cause an increase in TGF-β expression, particularly in renal
epithelial cells (Khanna et al., 1999a; Khanna et al., 1999b; McMorrow et al., 2005;
Shihab et al., 1996); these observations complement the overall immunosuppressive
effects of TGF-β.
There are numerous reports of viruses that induce TGF-β expression, including
cytomegalovirus, hepatitis B virus, hepatitis C virus, and HIV; the immunosuppressive
effects of TGF-β have obvious benefits for the virus during infection (Li et al., 2006;
Reed, 1999). However, there are also several reports of viruses that are regulated by the
signaling events of the TGF-β pathway. TGF-β treatment was shown to inhibit viral
RNA replication and protein expression from a hepatitis C virus replicon (Murata et al.,
25
2005). Overexpression of Smad3 and Smad4 proteins regulates the activity of HIV-1
LTR in astrocytes (Coyle-Rink et al., 2002). TGF-β1 can induce the reactivation of lytic
EBV replication from latently-infected epithelial cells (Fukuda et al., 2001) and B cells
(di Renzo et al., 1994; Fahmi et al., 2000). In addition, JCV early and late promoter
activities are upregulated by Smad3 and Smad4 overexpression (Enam et al., 2004), and
treatment of JCV-infected astrocytes with TGF-β stimulates viral replication
(Ravichandran et al., 2007). Therefore, TGF-β expression and regulation is highly
relevant for viral infections.
Summary and Chapter Outline
The human polyomavirus BKV is able to establish a persistent, subclinical
infection in nearly the entire population, which demonstrates its ability to at least partially
evade the cytotoxic immune response. In immunosuppressed patients, particularly KT
and BMT recipients, BKV can reactivate to a productive lytic infection, suggesting that
the loss of certain immune components in conjunction with other factors alleviates
repression of viral infection. We hypothesized that soluble immune factors, such as
cytokines, may play an important role in regulating BKV. Therefore, the decrease in
lymphocyte proliferation and activation during immunosuppression may promote BKV
reactivation, as a result of reduced cytokine expression.
Previously, our lab has developed a cell culture system for the study of BKV
infection in the natural host cell for reactivation, renal proximal tubule epithelial cells
(Low et al., 2004). In this system, the absence of immune surveillance mimics the
environment of an immunosuppressed patient and may allow the examination of
cytokine-mediated effects on BKV in a highly relevant host cell. In Chapter II, we
26
describe the inhibitory effect of IFN-γ on BKV replication and demonstrate that the
primary level of regulation is early gene transcription. In Chapter III, we report the
differential effects of TGF-β on early promoter activity. In contrast to IFN-γ-mediated
regulation, which is similar for three rearranged strains of BKV examined, TGF-βmediated effects vary, depending on the strain of virus and the cell type. In Chapter IV,
we begin a deeper investigation to analyze the mechanism of BKV promoter regulation,
specifically chromatin remodeling, in the context of IFN-γ restriction of replication and
during infection with archetype virus. In Chapter V, we will discuss the relevance of our
findings and the overall future directions of this project.
27
References
Allander, T., Andreasson, K., Gupta, S., Bjerkner, A., Bogdanovic, G., Persson, M. A.
A., Dalianis, T., Ramqvist, T., Andersson, B., 2007. Identification of a third
human polyomavirus. J Virol 81, 4130-4136.
Anderson, H. A., Chen, Y., Norkin, L. C., 1996. Bound simian virus 40 translocates to
caveolin enriched membrane domains, and its entry is inhibited by drugs that
selectively disrupt caveolae. Mol Biol Cell 7, 1825-1834.
Andrews, C. A., Shah, K. V., Daniel, R. W., Hirsch, H. H., Rubin, R. H., 1988. A
serological investigation of BK virus and JC virus infections in recipients of renal
allografts. J Infect Dis 158, 176-181.
Ank, N., West, H., Paludan, S. R., 2006. IFN-lambda: novel antiviral cytokines. J
Interferon Cytokine Res 26, 373-379.
Annes, J. P., Munger, J. S., Rifkin, D. B., 2003. Making sense of latent TGFbeta
activation. J Cell Sci 116, 217-224.
Arthur, R. R., Shah, K. V., Baust, S. J., Santos, G. W., Saral, R., 1986. Association of BK
viruria with hemorrhagic cystitis in recipients of bone marrow transplants. N Engl
J Med 315, 230-234.
Atwood, W. J., 2001. Cellular receptors for the polyomaviruses. In (K. Khalili, and G. L.
Stoner, Eds.),"Human polyomaviruses: molecular and clinical perspectives",
Wiley-Liss, Inc., New York, pp. 179-196.
Azzi, A., Fanci, R., Bosi, A., Ciappi, S., Zakrzewska, K., de Santis, R., Laszlo, D., Guidi,
S., Saccardi, R., Vannucchi, A. M., Longo, G., Rossi-Ferrini, P., 1994.
Monitoring of polyomavirus BK viruria in bone marrow transplantation patients
by DNA hybridization assay and by polymerase chain reaction: an approach to
assess the relationship between BK viruria and hemorrhagic cystitis. Bone
Marrow Transplant 14, 235-240.
Bedi, A., Miller, C. B., Hanson, J. L., Goodman, S., Ambinder, R. F., Charache, P.,
Arthur, R. R., Jones, R. J., 1995. Association of BK virus with failure of
prophylaxis against hemorrhagic cystitis following bone marrow transplantation.
J Clin Oncol 13, 1103-1109.
Bernhoff, E., Gutteberg, T. J., Sandvik, K., Hirsch, H. H., Rinaldo, C. H., 2008.
Cidofovir inhibits polyomavirus BK replication in human renal tubular cells
downstream of viral early gene expression. Am J Transplant 8, 1413-1422.
28
Binggeli, S., Egli, A., Schaub, S., Binet, I., Mayr, M., Steiger, J., Hirsch, H. H., 2007.
Polyomavirus BK-specific cellular immune response to VP1 and large T-antigen
in kidney transplant recipients. Am J Transplant 7, 1-9.
Biron, C. A., Sen, G. C., 2001. Interferons and other cytokines. Third ed. In (D. M.
Knipe, and P. M. Howley, Eds.),"Fields Virology". Vol. 1, Lippincott Williams &
Wilkins, Philadelphia, pp. 321-351.
Biron, C. A., Sen, G. C., 2007. Innate responses to viral infections. Fifth ed. In (D. M.
Knipe, and P. M. Howley, Eds.),"Fields Virology". Vol. 1, Lippincott Williams &
Wilkins, Philadelphia, pp. 249-278.
Bitzer, M., von Gersdorff, G., Liang, D., Dominguez-Rosales, A., Beg, A. A., Rojkind,
M., Bottinger, E. P., 2000. A mechanism of suppression of TGF-beta/SMAD
signaling by NF-kB/RelA. Genes Dev 14, 187-197.
Blobe, G. C., Schiemann, W. P., Lodish, H. F., 2000. Role of transforming growth factorbeta in human disease. N Engl J Med 342, 1350-1358.
Bohl, D. L., Storch, G. A., Ryschkewitsch, C., Gaudreault-Keener, M., Schnitzler, M. A.,
Major, E. O., Brennan, D. C., 2005. Donor origin of BK virus in renal
transplantation and role of HLA C7 in susceptibility to sustained BK viremia. Am
J Transplant 5, 2213-2221.
Brennan, D. C., Agha, I., Bohl, D. L., Schnitzler, M. A., Hardinger, K. L., Lockwood,
M., Torrence, S., Schuessler, R., Roby, T., Gaudreault-Keener, M., Storch, G. A.,
2005. Incidence of BK with tacrolimus versus cyclosporine and impact of
preemptive immunosuppression reduction. Am J Transplant 5, 582-594.
Brown, K. A., Pietenpol, J. A., Moses, H. L., 2007. A tale of two proteins: differential
roles and regulation of Smad2 and Smad3 in TGF-beta signaling. J Cell Biochem
101, 9-33.
Chakraborty, T., Das, G. C., 1989. Identification of HeLa cell nuclear factors that bind to
and activate the early promoter of human polyomavirus BK in vitro. Mol Cell Biol
9, 3821-3828.
Chatterjee, M., Weyandt, T. B., Frisque, R. J., 2000. Identification of archetype and
rearranged forms of BK virus in leukocytes from healthy individuals. J Med Virol
60, 353-362.
Chen, Y., Trofe, J., Gordon, J., Du Pasquier, R. A., Roy-Chaudhury, P., Kuroda, M. J.,
Woodle, E. S., Khalili, K., Koralnik, I. J., 2006. Interplay of cellular and humoral
immune responses against BK virus in kidney transplant recipients with
polyomavirus nephropathy. J Virol 80, 3495-3505.
29
Chesters, P. M., Heritage, J., McCance, D. J., 1983. Persistence of DNA sequences of BK
virus and JC virus in normal human tissues and in diseased tissues. J Infect Dis
147, 676-684.
Cicala, C., Avantaggiati, M. L., Graesmann, A., Rundell, K., Levine, A. S., Carbone, M.,
1994. Simian virus 40 small-t antigen stimulates viral DNA replication in
permissive monkey cells. J Virol 68, 3138-3144.
Co, J. K., Verma, S., Gurjav, U., Sumibcay, L., Nerurkar, V. R., 2007. Interferon-alpha
and -beta restrict polyomavirus JC replication in primary human fetal glial cells:
implications for progressive multifocal leukoencephalopathy therapy. J Infect Dis
196, 712-718.
Cole, C. N., Conzen, S. D., 2001. Polyomaviridae: the viruses and their replication.
Fourth ed. In (D. M. Knipe, and P. M. Howley, Eds.),"Fields Virology". Vol. 2,
Lippincott Williams &Wilkins, Philadelphia, pp. 2141-2174.
Coleman, D. V., Wolfendale, M. R., Daniel, R. A., Dhanjal, N. K., Gardner, S. D.,
Gibson, P. E., Field, A. M., 1980. A prospective study of human polyomavirus
infection in pregnancy. J Infect Dis 142, 1-8.
Comoli, P., Azzi, A., Maccario, R., Basso, S., Botti, G., Basile, G., Fontana, I., Labirio,
M., Cometa, A., Poli, F., Perfumo, F., Locatelli, F., Ginevri, F., 2004.
Polyomavirus BK-specific immunity after kidney transplantation. Transplantation
78, 1229-1232.
Comoli, P., Binggeli, S., Ginevri, F., Hirsch, H. H., 2006. Polyomavirus-associated
nephropathy: update on BK virus-specific immunity. Transpl Infect Dis 8, 86-94.
Cotterill, H. A., Macaulay, M. E., Wong, V., 1992. Reactivation of polyomavirus in bone
marrow transplant recipients. J Clin Pathol 45, 445.
Coyle-Rink, J., Sweet, T. M., Abraham, S., Sawaya, B. E., Batuman, O., Khalili, K.,
Amini, S., 2002. Interaction between TGFbeta signaling proteins and C/EBP
controls basal and Tat-mediated transcription of HIV-1 LTR in astrocytes.
Virology 299, 240-247.
Cubitt, C. L., 2006. Molecular genetics of the BK virus. Adv Exp Med Biol 577, 85-95.
De Mattei, M., Martini, F., Tognon, M., Serra, M., Baldini, N., Barbanti-Brodano, G.,
1994. Polyomavirus latency and human tumors. J Infect Dis 169, 1175-1176.
Del Vecchio, A. M., Steinman, R. A., Ricciardi, R. P., 1989. An element of the BK virus
enhancer required for DNA replication. J Virol 63, 1514-1524.
30
Der, S. D., Zhou, A., Williams, B. R., Silverman, R. H., 1998. Identification of genes
differentially regulated by interferon alpha, beta, or gamma using oligonucleotide
arrays. Proc Natl Acad Sci U S A 95, 15623-15628.
Derynck, R., Zhang, Y. E., 2003. Smad-dependent and Smad-independent pathways in
TGF-beta family signalling. Nature 425, 577-584.
Deyerle, K. L., Sajjadi, F. G., Subramani, S., 1989. Analysis of origin of DNA replication
of human papovavirus BK. J Virol 63, 356-365.
Deyerle, K. L., Subramani, S., 1988. Linker scan analysis of the early regulatory region
of human papovavirus BK. J Virol 62, 3378-3387.
di Renzo, L., Altiok, A., Klein, G., Klein, E., 1994. Endogenous TGF-beta contributes to
the induction of the EBV lytic cycle in two Burkitt lymphoma cell lines. Int J
Cancer 57, 914-919.
Doerries, K., Vogel, E., Gunther, S., Czub, S., 1994. Infection of human polyomaviruses
JC and BK in peripheral blood leukocytes from immunocompetent individuals.
Virology 198, 59-70.
Drachenberg, C. B., Hirsch, H. H., Ramos, E., Papadimitriou, J. C., 2005. Polyomavirus
disease in renal transplantation: review of pathological findings and diagnostic
methods. Hum Pathol 36, 1245-1255.
Dubois, C. M., Laprise, M. H., Blanchette, F., Gentry, L. E., Leduc, R., 1995. Processing
of transforming growth factor-beta1 precursor by human furin convertase. J Biol
Chem 270, 10618-10624.
Dugan, A. S., Eash, S., Atwood, W. J., 2005. An N-linked glycoprotein with alpha(2,3)linked sialic acid is a receptor for BK virus. J Virol 79, 14442-14445.
Eash, S., Atwood, W. J., 2005. Involvement of cytoskeletal components in BK virus
infectious entry. J Virol 79, 11734-11741.
Eash, S., Querbes, W., Atwood, W. J., 2004. Infection of Vero cells by BK virus is
dependent on caveolae. J Virol 78, 11583-11590.
Egli, A., Binggeli, S., Bodaghi, S., Dumoulin, A., Funk, G. A., Khanna, N., Leuenberger,
D., Gosert, R., Hirsch, H. H., 2007. Cytomegalovirus and polyomavirus BK
posttransplant. Nephrol Dial Transplant 22, viii72-viii82.
Elsner, C., Doerries, K., 1992. Evidence of human polyomavirus BK and JC infection in
normal brain tissue. Virology 191, 72-80.
31
Enam, S., Sweet, T. M., Amini, S., Khalili, K., Del Valle, L., 2004. Evidence for
involvement of transforming growth factor beta1 signaling pathway in activation
of JC virus in human immunodeficiency virus 1-associated progressive multifocal
leukoencephalopathy. Arch Pathol Lab Med 128, 282-291.
Fahmi, H., Cochet, C., Hmama, Z., Opolon, P., Joab, I., 2000. Transforming growth
factor beta 1 stimulates expression of the Epstein-Barr virus BZLF1 immediateearly gene product ZEBRA by an indirect mechanism which requires the MAPK
kinase pathway. J Virol 74, 5810-5818.
Feng, H., Shuda, M., Chang, Y., Moore, P. S., 2008. Clonal integration of a polyomavirus
in human Merkel cell carcinoma. Science 319, 1096-1100.
Feng, X. H., Derynck, R., 2005. Specificity and versatility in TGF-beta signaling through
Smads. Annu Rev Cell Dev Biol 21, 659-693.
Fukuda, M., Ikuta, K., Yanagihara, K., Tajima, M., Kuratsune, H., Kurata, T., Sairenji,
T., 2001. Effect of transforming growth factor-beta1 on the cell growth and
Epstein-Barr virus reactivation in EBV-infected epithelial cell lines. Virology 288,
109-118.
Funk, G. A., Steiger, J., Hirsch, H. H., 2006. Rapid dynamics of polyomavirus type BK
in renal transplant recipients. J Infect Dis 193, 80-87.
Gardner, S. D., Field, A. M., Coleman, D. V., Hulme, B., 1971. New human papovavirus
(B.K.) isolated from urine after renal transplantation. Lancet 1, 1253-1257.
Gaynor, A. M., Nissen, M. D., Whiley, D. M., Mackay, I. M., Lambert, S. B., Wu, G.,
Brennan, D. C., Storch, G. A., Sloots, T. P., Wang, D., 2007. Identification of a
novel polyomavirus from patients with acute respiratory tract infections. PLoS
Pathogens 3, 0595-0603.
Giannopoulou, M., Iszkula, S. C., Dai, C., Tan, X., Yang, J., Michalopoulos, G. K., Liu,
Y., 2006. Distinctive roles for Stat3 and Erk-1/2 activation in mediating
interferon-gamma inhibition of TGF-beta1 action. Am J Physiol Renal Physiol
290, F1234-F1240.
Gosert, R., Rinaldo, C. H., Funk, G. A., Egli, A., Ramos, E., Drachenberg, C. B., Hirsch,
H. H., 2008. Polyomavirus BK with rearranged noncoding control region emerge
in vivo in renal transplant patients and increase viral replication and
cytopathology. J Exp Med 205, 841-852.
Goudsmit, J., Wertheim-van Dillen, P., van Strien, A., van der Noordaa, J., 1982. The
role of BK virus in acute respiratory tract disease and the presence of BKV DNA
in tonsils. J Med Virol 10, 91-99.
32
Grinde, B., Gayorfar, M., Rinaldo, C. H., 2007. Impact of a polyomavirus (BKV)
infection on mRNA expression in human endothelial cells. Virus Res 123, 86-94.
Grinnell, B. W., Berg, D. T., Walls, J. D., 1988. Negative regulation of the human
polyomavirus BK enhancer involves cell-specific interaction with a nuclear
repressor. Mol Cell Biol 8, 3448-3457.
Harris, K. F., Christensen, J. B., Imperiale, M. J., 1996. BK virus large T antigen:
interactions with the retinoblastoma family of tumor suppressor proteins and
effects on cellular growth control. J Virol 70, 2378-2386.
Harris, K. F., Christensen, J. B., Radany, E. H., Imperiale, M. J., 1998. Novel
mechanisms of E2F induction by BK virus large-T antigen: requirement of both
the pRb-binding and the J domains. Mol Cell Biol 18, 1746-1756.
Hata, A., Lagna, G., Massague, J., Hemmati-Brivanlou, A., 1998. Smad6 inhibits
BMP/Smad1 signaling by specifically competing with the Smad4 tumor
suppressor. Genes Dev 12, 186-197.
Hayashi, H., Abdollah, S., Qiu, Y., Cai, J., Xu, Y. Y., Grinnell, B. W., Richardson, M.
A., Topper, J. N., Gimbrone, M. A., Wrana, J. L., Falb, D., 1997. The MADrelated protein Smad7 associates with the TGFbeta receptor and functions as an
antagonist of TGFbeta signaling. Cell 89, 1165-1173.
Heritage, J., Chesters, P. M., McCance, D. J., 1981. The persistence of papovavirus BK
DNA sequences in normal human renal tissue. J Med Virol 8, 143-150.
Hirsch, H. H., 2005. BK virus: opportunity makes a pathogen. Clin Infect Dis 41, 354360.
Hirsch, H. H., Drachenberg, C. B., Steiger, J., Ramos, E., 2006. Polyomavirus-associated
nephropathy in renal transplantation: critical issues of screening and management.
Adv Exp Med Biol 577, 160-173.
Hirsch, H. H., Knowles, W., Dickenmann, M., Passweg, J., Klimkait, T., Mihatsch, M. J.,
Steiger, J., 2002. Prospective study of polyomavirus type BK replication and
nephropathy in renal-transplant recipients. N Engl J Med 347, 488-496.
Howe, A. K., Gaillard, S., Bennett, J. S., Rundell, K., 1998. Cell cycle progression in
monkey cells expression simian virus 40 small t antigen from adenovirus vectors.
J Virol 72, 9637-9644.
Imperiale, M. J., 2001. The human polyomaviruses: an overview. In (K. Khalili, and G.
L. Stoner, Eds.),"Human polyomaviruses: molecular and clinical perspectives",
Wiley-Liss, Inc, New York, pp. 53-71.
33
Imperiale, M. J., Major, E. O., 2007. Polyomaviruses. Fifth ed. In (D. M. Knipe, and P.
M. Howley, Eds.),"Fields Virology". Vol. 2, Lippincott Williams & Wilkins,
Philadelphia, pp. 2263-2298.
Isaacs, A., Lindenmann, J., 1957. Virus interference. I. The interferon. Proc R Soc Lond
B Biol Sci 147, 258-267.
Jin, L., 1993. Rapid genomic typing of BK virus directly from clinical specimens. Mol
Cell Probes 7, 331-334.
Jin, L., Gibson, P. E., Booth, J. C., Clewley, J. P., 1993. Genomic typing of BK virus in
clinical specimens by direct sequencing of polymerase chain reaction products. J
Med Virol 41, 11-17.
Johnsen, J. I., Seternes, O. M., Johansen, T., Moens, U., Mantyjarvi, R., Traavik, T.,
1995. Subpopulations of non-coding control region variants within a cell culturepassaged stock of BK virus: sequence comparisons and biological characteristics.
J Gen Virol 76, 1571-1581.
Khanna, A., Cairns, V., Hosenpud, J. D., 1999a. Tacrolimus induces increased expression
of transforming growth factor-beta1 in mammalian lymphoid as well as
nonlymphoid cells. Transplantation 67, 614-619.
Khanna, A. K., Cairns, V. R., Becker, C. G., Hosenpud, J. D., 1999b. Transforming
growth factor (TGF)-beta mimics and anti-TGF-beta antibody abrogates the in
vivo effects of cyclosporine: demonstration of a direct role of TGF-beta in
immunosuppression and nephrotoxicity of cyclosporine. Transplantation 67, 882889.
Kierstad, T. D., Pipas, J. M., 1996. Database of mutations that alter the large tumor
antigen of simian virus 40. Nucleic Acids Res 24, 125-126.
Kim, H. S., Henson, J. W., Frisque, R. J., 2001. Transcription and replication in the
human polyomaviruses. In (K. Khalili, and G. L. Stoner, Eds.),"Human
polyomaviruses: molecular and clinical perspectives", Wiley-Liss, Inc, New
York, pp. 73-126.
Knepper, J. E., diMayorca, G., 1987. Cloning and characterization of BK virus-related
DNA sequences from normal and neoplastic tissues. J Med Virol 21, 289-299.
Knowles, W. A., Pipkin, P., Andrews, N., Vyse, A., Minor, P., Brown, D. W. G., Miller,
E., 2003. Population-based study of antibody to the human polyomaviruses BKV
and JCV and the simian polyomavirus SV40. J Med Virol 71, 115-123.
34
Li, M. O., Wan, Y. Y., Sanjabi, S., Robertson, A. K., Flavell, R. A., 2006. Transforming
growth factor-beta regulation of immune responses. Annu Rev Immunol 24, 99146.
Li, T. C., Takeda, N., Kato, K., Nilsson, J., Xing, L., Haag, L., Cheng, R. H., Miyamura,
T., 2003. Characterization of self-assembled virus-like particles of human
polyomavirus BK generated by recombinant baculoviruses. Virology 311, 115124.
Low, J., Humes, H. D., Szczypka, M., Imperiale, M., 2004. BKV and SV40 infection of
human kidney tubular epithelial cells in vitro. Virology 323, 182-188.
Low, J. A., Magnuson, B., Tsai, B., Imperiale, M. J., 2006. Identification of gangliosides
GD1b and GT1b as receptors for BKV virus. J Virol 80, 1361-1366.
Manfredi, J. J., Prives, C., 1994. The transforming activity of simian virus 40 large tumor
antigen. Biochim Biophys Acta 1198, 65-83.
Markowitz, R. B., Dynan, W. S., 1988. Binding of cellular proteins to the regulatory
region of BK virus DNA. J Virol 62, 3388-3398.
Markowitz, R. B., Eaton, B. A., Kubik, M. F., Latorra, D., McGregor, J. A., Dynan, W.
S., 1991. BK virus and JC virus shed during pregnancy have predominantly
archetypal regulatory regions. J Virol 65, 4515-4519.
Markowitz, R. B., Tolbert, S., Dynan, W. S., 1990. Promoter evolution in BK virus:
Functional elements are created at sequence junctions. J Virol 64, 2411-2415.
McMorrow, T., Gaffney, M. M., Slattery, C., Campbell, E., Ryan, M. P., 2005.
Cyclosporine A induced epithelial-mesenchymal transition in human renal
proximal tubular epithelial cells. Nephrol Dial Transplant 20, 2215-2225.
Meneguzzi, G., Pignatti, P. F., Barbanti-Brodano, G., Milanesi, G., 1978.
Minichromosome from BK virus as a template for transcription in vitro. Proc Natl
Acad Sci U S A 75, 1126-1130.
Mengel, M., Marwedel, M., Radermacher, J., Eden, G., Schwarz, A., Haller, H., Kreipe,
H., 2003. Incidence of polyomavirus-nephropathy in renal allografts: influence of
modern immunosuppressive drugs. Nephrol Dial Transplant 18, 1190-1196.
Moens, U., Rekvig, O. P., 2001. Molecular biology of BK virus and clinical and basic
aspects of BK virus renal infection. In (K. Khalili, and G. L. Stoner,
Eds.),"Human polyomaviruses: molecular and clinical perspectives", Wiley-Liss,
Inc., New York, pp. 359-408.
35
Moens, U., Van Ghelue, M., 2005. Polymorphism in the genome of non-passaged human
polyomavirus BK: implications for cell tropism and the pathological role of the
virus. Virology 331, 209-231.
Monini, P., Rotola, A., Di Luca, D., De Lellis, L., Chiari, E., Corallini, A., Cassai, E.,
1995. DNA rearrangements impairing BK virus productive infection in urinary
tract tumors. Virology 214, 273-279.
Moriyama, T., Sorokin, A., 2008. Intracellular trafficking pathway of BK virus in human
renal proximal tubular epithelial cells. Virology 371, 336-349.
Murata, T., Ohshima, T., Yamaji, M., Hosaka, M., Miyanari, Y., Hijikata, M.,
Shimotohno, K., 2005. Suppression of hepatitis C virus replicon by TGF-beta.
Virology 331, 407-417.
Nakao, A., Afrakhte, M., Moren, A., Nakayama, T., Christian, J. L., Heuchel, R., Itoh, S.,
Kawabata, M., Heldin, N. E., Heldin, C. H., ten Dijke, P., 1997. Identification of
Smad7, a TGFbeta-inducible antagonist of TGF-beta signaling. Nature 389, 631635.
Negrini, M., Sabbioni, S., Arthur, R. R., Castagnoli, A., Barbanti-Brodano, G., 1991.
Prevalence of the archetypal regulatory region and sequence polymorphisms in
nonpassaged BK virus variants. J Virol 65, 5092-5095.
Nickeleit, V., Singh, H. K., Mihatsch, M. J., 2003. Polyomavirus nephropathy:
morphology, pathophysiology, and clinical management. Curr Opin Nephrol
Hypertens 12, 599-605.
Nukuzuma, S., Takasaka, T., Zheng, H. Y., Zhong, S., Chen, Q., Kitamura, T., Yogo, Y.,
2006. Subtype I BK polyomavirus strains grow more efficiently in human renal
epithelial cells than subtype IV strains. J Gen Virol 87, 1893-1901.
Padgett, B. L., Walker, D. L., ZuRhein, G. M., Eckroade, R. J., Dessel, B. H., 1971.
Cultivation of a papova-like virus from human brain with progressive multifocal
leukoencephalopathy. Lancet i, 1257-1260.
Pavlakis, M., Haririan, A., Klassen, D. K., 2006. BK virus infection after non-renal
transplantation. Adv Exp Med Biol 577, 185-189.
Pestka, S., Krause, C. D., Walter, M. R., 2004. Interferons, interferon-like cytokines, and
their receptors. Immunol Rev 202, 8-32.
Pho, M. T., Ashok, A., Atwood, W. J., 2000. JC virus enters human glial cells by clathrin
dependent receptor mediated endocytosis. J Virol 74, 2288-2292.
36
Priftakis, P., Bogdanovic, G., Kalantari, M., Dalianis, T., 2001. Overrepresentation of
point mutations in the Sp1 site of the non-coding control region of BK virus in
bone marrow transplanted patients with haemorrhagic cystitis. J Clin Virol 21, 17.
Prosser, S. E., Orentas, R. J., Jurgens, L., Cohen, E. P., Hariharan, S., 2008. Recovery of
BK virus large T-antigen-specific cellular immune response correlates with
resolution of BK virus nephritis. Transplantation 85, 185-192.
Radhakrishnan, S., Otte, J., Enam, S., Del Valle, L., Khalili, K., Gordon, J., 2003. JC
virus-induced changes in cellular gene expression in primary human astrocytes. J
Virol 77, 10638-10644.
Rahimi, R. A., Leof, E. B., 2007. TGF-beta signaling: a tale of two responses. J Cell
Biochem 102, 593-608.
Ravichandran, V., Jensen, P. N., Major, E. O., 2007. MEK1/2 inhibitors block basal and
transforming growth factor beta1-stimulated JC virus multiplication. J Virol 81,
6412-6418.
Reed, S. G., 1999. TGF-beta in infections and infectious diseases. Microbes Infect 1,
1313-1325.
Rinaldo, C. H., Hansen, H., Traavik, T., 2005. Human endothelial cells allow passage of
an archetypal BK virus (BKV) strain-a tool for cultivation and functional studies
of natural BKV strains. Arch Virol 150, 1449-1458.
Rinaldo, C. H., Traavik, T., Hey, A., 1998. The agnogene of the human polyomavirus BK
is expressed. J Virol 72, 6233-6236.
Rubinstein, R., Pare, N., Harley, E. H., 1987. Structure and function of the transcriptional
control region of nonpassaged BK virus. J Virol 61, 1747-1750.
Rubinstein, R., Schoonakker, B. C., Harley, E. H., 1991. Recurring theme of changes in
the transcriptional control region of BK virus during adaptation to cell culture. J
Virol 65, 1600-1604.
Salzman, N. P., Natarajan, V., Selzer, G. B., 1986. Transcription of SV40 and polyoma
virus and its regulation. In (N. P. Salzman, Ed.),"The Papovaviridae", Plenum,
New York, pp. 27-98.
Sanda, C., Weitzel, P., Tsukahara, T., Schaley, J., Edenberg, H. J., Stephens, M. A.,
McClintick, J. N., Blatt, L. M., Li, L., Brodsky, L., Taylor, M. W., 2006.
Differential gene induction by type I and type II interferons and their
combination. J Interferon Cytokine Res 26, 462-472.
37
Seif, I., Khoury, G., Dhar, R., 1979. The genome of the human papovavirus BKV. Cell
18, 963-977.
Sen, G. C., 2001. Viruses and interferons. Annu Rev Microbiol 55, 255-281.
Shadan, F. F., Villarreal, L. P., 1993. Coevolution of persistently infecting small DNA
viruses and their hosts linked to host-interactive regulatory domains. Proc Natl
Acad Sci U S A 90, 4117-4121.
Sharma, P. M., Gupta, G., Vats, A., Shapiro, R., Randhawa, P. S., 2007. Polyomavirus
BK non-coding control region rearrangements in health and disease. J Med Virol
79, 1199-1207.
Shi, Y., Massague, J., 2003. Mechanisms of TGF-beta signaling from the cell membrane
to the nucleus. Cell 113, 685-700.
Shi, Y., Wang, Y. F., Jayaraman, L., Yang, H., Massague, J., Pavletich, N. P., 1998.
Crystal structure of a Smad MH1 domain bound to DNA: insights on DNA
binding in TGF-beta signaling. Cell 94, 585-594.
Shihab, F. S., Andoh, T. F., Tanner, A. M., Noble, N. A., Border, W. A., Franceschini,
N., Bennett, W. M., 1996. Role of transforming growth factor-beta 1 in
experimental chronic cyclosporine nephropathy. Kidney Int 49, 1141-1151.
Shivakumar, C. V., Das, G. C., 1996. Interaction of human polyomavirus BK with the
tumor-suppressor protein p53. Oncogene 13, 323-332.
Singh, H. K., Bubendorf, L., Mihatsch, M. J., Drachenberg, C. B., Nickeleit, V., 2006.
Urine cytology findings of polyomavirus infections. Adv Exp Med Biol 577, 201212.
Skoczylas, C., Fahrbach, K. M., Rundell, K., 2004. Cellular targets of the SV40 small-t
antigen in human cell transformation. Cell Cycle 3, 606-610.
Stahl, H., Droge, P., Knippers, R., 1986. DNA helicase activity of SV40 large tumor
antigen. EMBO J 5, 1939-1944.
Stillman, B., Gerard, R. D., Guggenheimer, R. A., Gluzman, Y., 1985. T antigen and
template requirements for SV40 DNA replication in vitro. EMBO J 4, 2933-2939.
Sugimoto, C., Hara, K., Taguchi, F., Yogo, Y., 1989. Growth efficiency of naturally
occurring BK virus variants in vivo and in vitro. J Virol 63, 3195-3199.
Sundsfjord, A., Flaegstad, T., Flo, R., Spein, A. R., Pedersen, M., Permin, H., Julsrud, J.,
Traavik, T., 1994. BK and JC viruses in human immunodeficiency virus type 1-
38
infected persons: prevalence, excretion, viremia, and viral regulatory regions. J
Infect Dis 169, 485-490.
Sundsfjord, A., Johansen, T., Flaegstad, T., Moens, U., Villand, P., Subramani, S.,
Traavik, T., 1990. At least two types of control regions can be found among
naturally occurring BK virus strains. J Virol 64, 3864-3871.
Sundsfjord, A., Osei, A., Rosenqvist, H., Van Ghelue, M., Silsand, Y., Haga, H. J.,
Rekvig, O. P., Moens, U., 1999. BK and JC viruses in patients with systemic
lupus erythematosus: prevalent and persistent BK viruria, sequence stability of the
viral regulatory regions, and nondetectable viremia. J Infect Dis 180, 1-9.
Swaminathan, S., Rajan, P., Savinova, O., Jagus, R., Thimmapaya, B., 1996. Simian virus
40 large-T bypasses the translational block imposed by the phosphorylation of
eIF-2alpha. Virology 219, 321-323.
Takasaka, T., Goya, N., Tokumoto, T., Tanabe, K., Toma, H., Ogawa, Y., Hokama, S.,
Momose, A., Funyu, T., Fujioka, T., Omori, S., Akiyama, H., Chen, Q., Zheng, H.
Y., Ohta, N., Kitamura, T., Yogo, Y., 2004. Subtypes of BK virus prevalent in
Japan and variation in their transcriptional control region. J Gen Virol 85, 28212827.
Tan, H., Derrick, J., Hong, J., Sanda, C., Grosse, W. M., Edenberg, H. J., Taylor, M. W.,
Seiwert, S., Blatt, L. M., 2005. Global transcriptional profiling demonstrates the
combination of type I and type II interferon enhances antiviral and immune
responses at clinically relevant doses. J Interferon Cytokine Res 25, 632-649.
ter Schegget, J., Sol, C. J. A., Wouters Baan, E., van der Noordaa, J., van Ormondt, H.,
1985. Naturally occurring BK virus variants (JL and Dik) with deletions in the
putative early enhancer-promoter sequences. J Virol 53, 302-305.
Trowbridge, P. W., Frisque, R. J., 1995. Identification of three new JC virus proteins
generated by alternative splicing of the early viral mRNA. J Neurovirol 1, 195206.
Ulloa, L., Doody, J., Massague, J., 1999. Inhibition of transforming growth factorbeta/SMAD signalling by the interferon-gamma/STAT pathway. Nature 397, 710713.
van Boxel-Dezaire, A. H. H., Stark, G. R., 2007. Cell type-specific signaling in response
to interferon-gamma. Curr Top Microbiol Immunol 316, 119-154.
Vats, A., Randhawa, P., Shapiro, R., 2006. Diagnosis and treatment of BK virusassociated transplant nephropathy. Adv Exp Med Biol 577, 213-227.
39
Verma, S., Ziegler, K., Ananthula, P., Co, J. K., Frisque, R. J., Yanagihara, R., Nerurkar,
V. R., 2006. JC virus induces altered patterns of cellular gene expression:
interferon-inducible genes as major transcriptional targets. Virology 345, 457-467.
Watanabe, S., Yoshiike, K., 1982. Change of DNA near the origin of replication
enhances the transforming capacity of human papovavirus BK. J Virol 42, 978985.
Watanabe, S., Yoshiike, K., 1986. Evolutionary changes of transcriptional control region
in a minute-plaque viable deletion mutant of BK virus. J Virol 59, 260-266.
Weihua, X., Ramanujam, S., Lindner, D. J., Kudaravalli, R. D., Freund, R., Kalvakolanu,
D. V., 1998. The polyoma virus T antigen interferes with interferon-inducible
gene expression. Proc Natl Acad Sci U S A 95, 1085-1090.
Whalen, B., Laffin, J., Friedrich, T. D., Lehman, J. M., 1999. SV40 small t antigen
enhances progression to >G2 during lytic infection. Exp Cell Research 251, 121127.
Wheelock, E. F., 1965. Interferon-like virus-inhibitor induced in human leukocytes by
phytohemagglutinin. Science 149, 310-311.
Yang, S. I., Lickteig, R. L., Estes, R., Rundell, K., Walter, G., Mumby, M. C., 1991.
Control of protein phosphatase 2A by simian virus 40 small-t antigen. Mol Cell
Biol 11, 1988-1995.
Yang, Z., Mu, Z., Dabovic, B., Jurukovski, V., Yu, D., Sung, J., Xiong, X., Munger, J. S.,
2007. Absence of integrin-mediated TGFbeta1 activation in vivo recapitulates the
phenotype of TGFbeta1-null mice. J Cell Biol 176, 787-793.
Zawel, L., Dai, J. L., Buckhaults, P., Zhou, S., Kinzler, K. W., Vogelstein, B., Kern, S.
E., 1998. Human Smad3 and Smad4 are sequence-specific transcription
activators. Mol Cell 1, 611-617.
Zerrahn, J., Knippschild, U., Winkler, T., Deppert, W., 1993. Independent expression of
the transforming amino-terminal domain of SV40 large T antigen from an
alternatively spliced third SV40 early mRNA. EMBO J 12, 4739-4746.
Zhong, S., Zheng, H. Y., Suzuki, M., 2007. Age-related urinary excretion of BK
polyomavirus by non-immunocompromised individuals. J Clin Microbiol 45,
193-198.
40
CHAPTER II
INHIBITORY EFFECT OF INTERFERON-GAMMA ON BK VIRUS GENE
EXPRESSION AND REPLICATION
Polyomavirus nephropathy (PVN) results in renal dysfunction and graft loss in up
to 10% of all kidney transplant recipients (Hirsch et al., 2005). It is widely accepted that
BK virus (BKV) is the etiological agent responsible for the majority of cases of PVN,
which are typically diagnosed within the first year after transplantation (Hirsch et al.,
2002; Hirsch and Steiger, 2003). PVN is characterized by the lytic, destructive
replication of BKV in proximal tubule epithelial cells in the transplanted kidney and is
normally diagnosed by renal biopsy to assess histological effects of infection, PCR to
determine viral presence and loads in the urine and blood, and the detection of decoy
cells, which are cells with distinct intranuclear inclusion bodies that are shed during
active BKV replication, in the urine (Drachenberg et al., 2005; Hirsch, 2005; Nickeleit et
al., 2003). Since there are currently no effective antiviral treatments for BKV infection,
the most common approach used to control PVN is to decrease the patient’s
immunosuppressive regimen. However, such an approach increases the risk of graft
rejection and thus is not an appealing strategy. The prevalence of PVN is increasing with
the advent of new, more powerful immunosuppressive therapies, making it a growing
concern for the transplant community.
41
The human polyomavirus BKV was first isolated in 1971 in the urine of a renal
transplant recipient (Gardner et al., 1971). BKV virions are small (40 to 45 nm in
diameter), non-enveloped, icosahedral, and contain a circular, double-stranded DNA
genome of approximately 5.2 kb (Moens and Rekvig, 2001), which is associated with
cellular histones to form a chromatin-like structure (Meneguzzi et al., 1978). The
genome encodes only six known proteins: the early proteins, large tumor antigen (TAg)
and small tumor antigen (tAg), and the late proteins, VP1, VP2, VP3, and agnoprotein.
BKV infects nearly the entire population, with seroprevalence reaching 60 to 80% by the
age of 10 (Knowles, 2001). BKV is thought to be contracted by respiratory transmission
and the primary infection is typically subclinical. Following the initial infection, BKV
spreads to other cells of the body, most notably peripheral blood mononuclear cells
(Doerries et al., 1994) and cells of the kidney and urinary tract (Chesters et al., 1983;
Heritage et al., 1981; Shinohara et al., 1993), in which the virus establishes a persistent,
subclinical infection. It is from these sites in immunocompromised patients that BKV
reactivates to a lytic infection, resulting in BKV-associated diseases such as PVN.
Previously, we described an in vitro system that allows the study of BKV lytic
infection of primary human proximal tubule epithelial cells (Low et al., 2004). The
functions of proximal tubule cells in the kidney include facilitating the recovery of blood
products, maintenance of blood pressure and volume, and production and release of
cytokines and chemokines to communicate with the host immune system (Briggs et al.,
2001; Daha and van Kooten, 2000). Proximal tubule cells remain in a differentiated state
for up to six passages in tissue culture (Humes et al., 2002) and thus provide an
environment similar to that which BKV encounters in an immunocompromised host. By
42
introducing individual elements of the immune system to this model, we can begin to
determine which components regulate BKV replication. In part because proximal tubule
epithelial cells are known to interact both with neighboring cells and with the immune
system through the production of cytokines and chemokines, namely IL-6, IL-8, IL-15,
TNF-α, MCP-1, RANTES, and TGF-β (Daha and van Kooten, 2000), we began to
investigate the role of cytokines in mediating regulation of BKV replication.
Our interest in the effects of cytokines on BKV replication was initiated by
clinical reports and observations that argued for the importance of the cell-mediated
immune response, as opposed to other arms of the immune system, in controlling BKV
persistent infection. First, it has been demonstrated that up to 90% of the adult
population has BKV-specific antibodies (Knowles, 2001) and patients with levels of antiBKV antibodies similar to those of healthy individuals still develop PVN (Hariharan et
al., 2005; Hirsch et al., 2002). Several reports detail the activation of the humoral
immune response in patients with PVN, although antibody levels are shown to increase
only after the stabilization of renal function and a decrease in viral load (Chen et al.,
2006; Comoli et al., 2004; Hariharan et al., 2005). Furthermore, high titers of BKVspecific antibodies in the donor correlate well with the incidence of BKV reactivation in
the recipient (Bohl et al., 2005). These findings suggest that the presence of BKVspecific antibodies does not prevent reactivation and development of PVN, although they
may be effective at controlling viremia (Chen et al., 2006). In addition, the failure of the
cytotoxic T lymphocyte (CTL) response to eliminate all cells harboring BKV, as
demonstrated by the well-documented periodic shedding of virus in immunocompetent
individuals, indicates that another facet of the immune system plays a role in controlling
43
persistent infection. There are several studies demonstrating the correlation of low levels
of BKV-specific interferon-gamma (IFN-γ)-producing cells and the development of
viremia and PVN (Chen et al., 2006; Comoli et al., 2004). Furthermore, others have
reported the inhibitory effect of IFN-γ on the promoters of SV40, a related polyomavirus,
and human cytomegalovirus, which lacks structural relatedness to BKV but has similar
epidemiological features, such as latency in the kidney and reactivation upon
immunosuppression (Harms and Splitter, 1995; Qin et al., 1997; Ritter et al., 2000).
These findings prompted us to investigate the potential regulation of BKV replication by
IFN-γ.
IFN-γ is the sole member of the type II family of interferons and is a secreted
glycoprotein of ~25 kDa produced primarily by natural killer cells during the innate
immune response and specific antigen activated-T lymphocytes during the adaptive
immune response (Pestka et al., 2004; Schroder et al., 2004). Activation of the IFNγ cascade within a cell is initiated by the binding of IFN-γ to the cell surface receptor and
subsequent activation of the JAK-STAT signaling pathway. JAK1 and JAK2 kinases,
which are associated with the two chains of the IFN-γ receptor, IFNGR1 and IFNGR2,
respectively, become phosphorylated upon IFN-γ binding. As a result, STAT1
homodimers are phosphorylated and subsequently translocated to the nucleus, where they
act as transcription factors to mediate regulation of IFN-γ responsive genes (Pestka et al.,
2004; Schroder et al., 2004). IFN-γ was first discovered by its antiviral activity
(Wheelock, 1965) and remains widely known as a potent antiviral cytokine. Among the
genes up-regulated by the IFN-γ signaling cascade are protein kinase R, which inhibits
cellular translation by phosphorylating and inactivating eukaryotic initiation factor-2; 2'44
5' oligoadenylate synthetase, which activates RNaseL to non-specifically degrade mRNA
transcripts and inhibit gene expression; and interferon regulatory factor 1 (IRF-1), which
up-regulates production of caspase 1 and promotes apoptosis of the cell. IFN-γ also plays
a major role in the activation and recruitment of immune cells and up-regulates the
expression of MHC class I and II molecules on the cell surface. More generally, IFNγ significantly (by at least two-fold) affects the expression of more than 100 genes in
mammalian cells (Der et al., 1998). In addition, other genes may be regulated indirectly
by a group of IFN-γ-responsive transcription factors that perpetuate the signaling
cascade. These transcription factors may also act on viral promoters to inhibit viral gene
expression and replication.
In this report, we characterize the specific inhibitory effect of IFN-γ on BKV gene
expression and replication during lytic infection of proximal tubule cells. We were
interested in determining the points at which the viral life cycle is affected and the
conditions in which IFN-γ has the strongest inhibitory effect on viral replication. IFN-γ
inhibited BKV gene expression, both at the level of transcription and translation, and
reduced the level of viral progeny produced during lytic infection. These results are
important for understanding the host immune response to BKV and, more specifically,
the role of cytokines in regulating BKV replication and infection.
45
Materials and Methods
Cell culture. Primary human renal proximal tubule epithelial cells (RPTE cells,
Cambrex) were maintained for up to six passages in renal epithelial cell basal medium
(REBM, Cambrex) supplemented with human epidermal growth factor, fetal bovine
serum, hydrocortisone, epinephrine, insulin, triiodothyronine, transferrin, and GA-1000
as indicated for renal epithelial cell growth medium (REGM, supplements obtained as
REGM SingleQuots, Cambrex). RPTE cells were grown at 37°C with 5% CO2 in a
humidified incubator.
Viruses. The genome of the TU strain of BKV was cloned into the EcoRI site of pGEM7Zf(-). Genomes of the Dunlop and Proto-2 strains of BKV were cloned into the BamHI
site of pBR322 (gift of P. M. Howley). BKV stocks were prepared from these genomic
clones: 4 μg plasmid DNA was digested with restriction enzymes (EcoRI for BKV(TU),
BamHI for BKV(Dunlop, Proto-2)), recircularized with T4 DNA ligase and phenolchloroform extracted, and the resulting DNA was transfected into one T75 flask of 60%
confluent RPTE cells (~ 4 x 106 cells) using Effectene (Qiagen). After three weeks, cells
and supernatants were collected and viral lysates were prepared by three freeze (-80°C) /
thaw (37°C) cycles. The resulting lysates were used to infect four T75 flasks of 70%
confluent RPTE cells and after three weeks, viral lysates were prepared as above. The
resulting viral stocks were titrated by fluorescent focus assay and the integrity of the noncoding control region (NCCR) was confirmed by sequencing of PCR products.
Cytokines and chemokines. Recombinant human IFN-γ, IL-6, IL-8, MCP-1, RANTES,
and TNF-α were purchased from PeproTech, Inc. and reconstituted according to
manufacturer’s recommendations. Recombinant human IFN-α was purchased from
46
Sigma and supplied as a solution in PBS. The doses used to treat BKV-infected RPTE
cells were as follows: IL-6, IL-8, MCP-1, TNF-α were used at 100 ng/ml, RANTES was
used at 300 ng/ml, IFN-α was used at 50 or 250 U/ml, and IFN-γ was used primarily at
50 or 250 U/ml, with the exception of the dose response experiment in which six fivefold dilutions were used, starting from 1250 U/ml.
Infections. Unless otherwise stated, 70% confluent RPTE cells were infected with the
TU strain of BKV at an MOI of 0.5, incubating for one hour at 37°C. Viral lysate used
for the infection was replaced with fresh media (REGM) and cytokines were added three
to six hours post-infection.
Western blotting. Unless otherwise stated, total cell protein was harvested at four days
post-infection using E1A lysis buffer (Harlow et al., 1986) supplemented with 5 μg /ml
PMSF, 5 μg /ml aprotinin, 5 μg /ml leupeptin, 0.05 M sodium fluoride, and 0.2 mM
sodium orthovanadate. The Bio-Rad protein assay was used to determine the protein
concentration of each lysate and 8 μg of protein were electrophoresed on an 8% SDSpolyacrylamide gel. Proteins were transferred to nitrocellulose membrane in 1x Towbin
buffer (25 mM Tris, 192 mM glycine, 20% methanol) for 12 to 14 hours at 60 volts
(constant) and 4°C. The following primary antibodies were diluted in PBS containing
0.1% Tween (PBS-T) and 5% nonfat dry milk: pAb416 (Harlow et al., 1981) for
detection of TAg expression, P5G6 (gift of D. Galloway) for detection of VP1
expression, and Ab8245 (Abcam) for detection of glyceraldehyde-3-phosphatedehydrogenase (GAPDH) expression as a loading control. Blots were washed in PBS-T,
probed with horseradish peroxidase-conjugated anti-mouse IgG secondary antibody
(Sigma), and developed using ECL+ reagent (Amersham) and exposure to film.
47
Fluorescent Focus Assay (FFA). Viral lysates were harvested at T = 0 (after one hour
absorption at 37°C) and four days post-infection (or at T = 0 and 1, 2, 3, 4, and 6 days
post-infection for the viral growth curve) by collecting and preparing infected cells and
supernatants by three freeze/thaw cycles, as described above. Seventy percent confluent
RPTE cells in 24-well plates were infected with 10-fold dilutions of viral lysates for four
days at 37°C. Cells were fixed with 50% methanol/50% acetone for 10 min at room
temperature, air dried for 15 min, wrapped in parafilm and stored overnight at -20°C, as
an antigen retrieval step. Plates were thawed briefly at room temperature, rehydrated
with PBS, and incubated for one hour at 37°C with pAb416 in PBS, followed by FITCconjugated anti-mouse IgG secondary antibody (Sigma) in PBS with 0.005% Evans Blue
stain. Titer was determined by counting five random fields in at least three replicate
wells and is expressed as infectious units per ml (IU/ml). Statistical significance was
determined using a two-tailed Student’s t test assuming unequal variance and P values
< 0.05 were considered significant.
RNA extraction and cDNA synthesis. Total cell RNA was harvested at two or four
days post-infection using TRIzol reagent (Invitrogen) according to manufacturer’s
instructions. RNA samples were treated with DNaseI (Promega) to reduce contaminating
DNA and the integrity of the RNA was confirmed by electrophoresis on an agarose gel.
To generate cDNA, a reverse transcription reaction was performed using 1 μg RNA as
template and the iScript cDNA Synthesis Kit (Bio-Rad), according to manufacturer’s
instructions.
Real Time PCR: Taqman Assay. Primers and probes were designed using Primer3
software (Rozen and Skaletsky, 2000) to amplify 90- and 105-base pair fragments of the
48
TAg and GAPDH genes, respectively; sequences are shown in Table 1. Primers were
synthesized by Invitrogen and probes, tagged with 6-fluorescein (FAM) as the reporter
dye at the 5' end and TAMRA-Sp as the quencher dye at the 3' end, were synthesized by
Integrated DNA Technologies, Inc. (IDT). Reactions were performed in a total volume
of 25 μl using TaqMan Universal PCR 2x master mix (Applied Biosystems), 2.5 μl
cDNA template, 500 nM of each primer, and 200 nM probe. Amplification was
performed in 96-well PCR plates (Bio-Rad) using the iCycler iQ5 Real Time Detection
System (Bio-Rad) with the following PCR program: 2 min at 50°C; 10 min at 95°C; 40
cycles of denaturation at 95°C for 15 sec and annealing and extension at 56°C for 1 min.
Results are presented as the fold change in TAg transcript levels, with the levels in
samples treated with 250 U/ml IFN-γ arbitrarily set to one. Results were normalized to
the levels of GAPDH transcripts present using the 2-ΔΔC(T) (Livak) method (Livak and
Schmittgen, 2001). Statistical significance was determined using a two-tailed Student’s t
test assuming unequal variance and P values < 0.05 were considered significant.
49
Gene
TAg
Sequence (5' to 3')
Forward Primer
AAAAATGGAGCAGGATGTAAAGGT
Reverse Primer
TCTTCTGTTCCATAGGTTGGCA
Probe
GAPDH
AGCTACTCCAGGTTCCAAAATCAGGCTGA
Forward Primer
GCCTCAAGATCATCAGCAAT
Reverse Primer
CTGTGGTCATGAGTCCTTCC
Probe
AAGGTCATCCATGACAACTTTGGTATCG
Table 1. Sequences of primers and probes used in Taqman Real Time PCR assays.
50
Results
IFN-γ inhibits the expression of viral genes during infection with BKV. To begin our
investigation of the role of cytokines and chemokines in regulating BKV infection, we
examined the effect of several molecules on viral gene expression during infection of
RPTE cells. The cytokines and chemokines used in this experiment were chosen for their
well-known antiviral effects (IFN-α and IFN-γ) or because they are produced by
proximal tubule epithelial cells following injury or stimulation (Daha and van Kooten,
2000). RPTE cells were infected with BKV, treated with IFN-γ, IFN-α, IL-6, IL-8,
MCP-1, RANTES, or TNF-α, and total cell protein was harvested at four days postinfection. As a control, the ability of each cytokine to stimulate RPTE cells was analyzed
by Western blot: phosphorylation of STAT1 was observed for IFN-γ and IFN-α, while
phosphorylation of ERK1/2 was observed for IL-6, IL-8, MCP-1, and TNF-α; results for
RANTES were inconclusive (data not shown). Western blot analysis of lysates probing
for the viral proteins TAg, representing early gene expression, and VP1, representing late
gene expression, revealed that only IFN-γ had a significant effect on viral gene
expression (Figure 2.1A). Similar levels of GAPDH, a cellular housekeeping gene, were
detected in all samples, showing that the inhibitory effect on viral gene expression was
specific and could not be attributed to the known non-specific cellular effects of IFN-γ,
namely the induction of anti-proliferative and pro-apoptotic pathways (Figure 2.1A).
Comparable results were obtained when samples were normalized to the total number of
cells per lysate (data not shown). Although IFN-α and IL-6 both appeared to have a slight
inhibitory effect on TAg and VP1, the potential roles of these cytokines in regulating
51
Figure 2.1. Dose-dependent IFN-γ inhibition of BKV gene expression. RPTE cells
were infected with the TU strain of BKV at an MOI of 0.5 and treated with cytokines at
three to six hours post-infection, and total cell protein was harvested at four days postinfection. Samples were analyzed by Western blotting, probing for TAg, VP1, and
GAPDH. A) Infected cells were treated with 50 or 250 U/ml IFN-γ or IFN-α, 100 ng/ml
IL-6, IL-8, MCP-1, or TNF-α, or 300 ng/ml RANTES. B) Infected cells were treated
with the indicated concentrations of IFN-γ. Mock, Mock-infected samples with no
cytokine treatment; untreated, BKV-infected samples with no cytokine treatment; T = 0,
samples harvested directly after one hour of adsorption with BKV.
52
BKV replication was not pursued at this time. The dramatic inhibition of viral protein
expression with IFN-γ treatment is the subject of further investigation in this study.
To determine whether IFN-γ exerts an inhibitory effect on BKV gene expression
in a dose-dependent manner, RPTE cells were infected with BKV and treated with fivefold dilutions of IFN-γ. Total cell protein was harvested at four days post-infection and
analyzed for TAg, VP1, and GAPDH expression by Western blotting (Figure 2.1B). At
the three higher doses (1250, 250, and 50 U/ml IFN-γ), levels of TAg were undetectable
while at the three lower doses (10, 2, and 0.4 U/ml IFN-γ), TAg expression approached
the level of the untreated sample. At higher doses, VP1 expression was detectable but
greatly reduced when compared to the untreated sample, and levels of VP1 increased as
the dose of IFN-γ decreased. Samples harvested directly after the initial infection (one
hour incubation with viral lysates), denoted T = 0, were analyzed to determine the
amount of detectable VP1 due to input virions, which was negligible (Figure 2.1B).
Thus, levels of VP1 in all samples corresponded to de novo protein expression during
infection. Levels of GAPDH were similar for all samples, indicating equal protein
loading and accounting for the general cellular effects of IFN-γ.
IFN-γ does not affect the kinetics of BKV replication. To determine whether IFN-γmediated inhibition of viral gene expression was the result of a delay in the progression
of infection or a reduction in the level of gene expression, we infected RPTE cells with
BKV, treated with 50 or 250 U/ml IFN-γ, and harvested total cell protein at 12 hour
intervals over four days (Figure 2.2A). Samples were assayed for TAg, VP1, and
GAPDH expression by Western blotting. In both untreated samples and samples treated
with 50 U/ml IFN-γ, TAg expression was first detected at 36 hours post-infection, while
53
VP1 expression was first detected at 48 hours post-infection, indicating that there was no
delay in the initiation of viral gene expression. Similar results were obtained from
analysis of samples treated with 250 U/ml IFN-γ (data not shown). An interesting pattern
of viral gene expression was apparent in treated samples: the levels of TAg and VP1
peaked at 72 and 84 hours post-infection, respectively, and then decreased at 96 hours
post-infection (Figure 2.2A). This level of inhibition of viral gene expression was
maintained up to 13 days post-infection (data not shown).
To examine the effect of IFN-γ on the timing and level of viral progeny produced,
viral lysates were harvested at T = 0 and 1, 2, 3, 4, and 6 days post-infection from BKVinfected RPTE cells either untreated or treated with 50 or 250 U/ml IFN-γ. The titer of
each lysate was determined by fluorescent focus assay, which detects the expression of
TAg in newly infected cells (Figure 2.2B). The increase in viral titer of untreated and
IFN-γ-treated samples was similar during the first 48 hours of infection, after which the
untreated samples abruptly increased in viral titer whereas the IFN-γ-treated samples
maintained the slower rate of increase seen in the first 48 hours. Interestingly, the viral
titers of both untreated and IFN-γ-treated samples reached a plateau at four days postinfection; however, the titers of the IFN-γ-treated samples were approximately 16- and
20-fold lower (for 50 and 250 U/ml IFN-γ-treated samples, respectively) than the
untreated samples at six days post-infection. The timing of viral protein expression and
progeny production were similar for untreated samples and samples treated with 50 or
250 U/ml IFN-γ, and comparable results were obtained when cells were treated with
IFN-γ prior to infection (data not shown). These data suggest that IFN-γ inhibits the level
of viral gene expression and does not delay the progression of infection. This finding
54
Figure 2.2. BKV replication kinetics in the presence of IFN-γ. RPTE cells were
infected with the TU strain of BKV at an MOI of 0.5, treated with 50 (Figure 2.2A, 2.2B)
or 250 U/ml (Figure 2.2B) IFN-γ at three to six hours post-infection. A) Total cell
protein was harvested every 12 hours for four days. Samples were analyzed by Western
blot, probing for TAg, VP1, and GAPDH. Untreated, BKV-infected samples with no
IFN-γ treatment; M, mock-infected samples; hpi, hours post-infection; 0 hpi, samples
harvested directly after one hour absorption with BKV (to detect protein from input
virions). B) Viral lysates were harvested at 0, 1, 2, 3, 4, and 6 days post-infection and
progeny production was determined by the fluorescent focus assay. Data are represented
as the log of the viral titer in infectious units per ml (IU/ml) and samples were assayed in
triplicate. Error bars are too small to be seen for some samples.
55
rules out a block at entry or trafficking to the nucleus as the mechanism of IFN-γ
inhibition.
Treatment with IFN-γ results in a reduction in the level of early region transcripts.
The previous experiments demonstrated the inhibitory effect of IFN-γ on the level of viral
protein expression and progeny production. To determine whether IFN-γ-mediated
inhibition occurs at the level of transcription or translation, we examined the effect of
IFN-γ on viral transcript production. Total cell RNA was harvested at 48 or 96 hours
post-infection from RPTE cells infected with BKV and treated with 50 or 250 U/ml
IFN-γ. RNA samples were analyzed by real time reverse-transcription PCR (real time
RT-PCR) to detect the levels of TAg transcripts. By designing the TAg primers and
probe set to amplify a 90-base pair amplicon across the splice site of the early region,
amplification was limited to the amplicon produced from the TAg cDNA template, as
opposed to that from the tAg or unspliced cDNA template. Results were normalized to
the levels of GAPDH transcripts to account for the non-specific cellular effects of IFN-γ,
using the 2-ΔΔC(T) method and are presented as the fold change in TAg transcript levels,
with the levels in samples treated with 250 U/ml IFN-γ arbitrarily set to one.
At 48 hours post-infection, there was a modest 1.6-fold decrease in TAg transcript
levels with 50 and 250 U/ml IFN-γ treatment (Figure 2.3A). These results were
confirmed by Northern blot analysis, using a probe specific for the entire BKV early
region to detect levels of all species of early region transcripts (data not shown). A
similar level of inhibition was observed at 48 hours post-infection at the level of protein
expression (Figure 2.2A). At 96 hours post-infection, there was a more dramatic and
highly significant inhibitory effect with IFN-γ treatment, such that treatments with 50 and
56
Figure 2.3. Effect of IFN-γ on viral early region transcript levels. RPTE cells were
infected with the TU strain of BKV at an MOI of 0.5, treated with 50 or 250 U/ml IFN-γ
at three to six hours post-infection, and total cell RNA was prepared at (A) 48 or (B) 96
hours post-infection. The level of TAg transcripts in each sample was determined using
real time RT-PCR, normalizing to level of GAPDH transcripts; samples treated with 250
U/ml IFN-γ were arbitrarily set to one. Each bar represents the average of two (Figure
2.3A) or three (Figure 2.3B) independent experiments analyzed in triplicate in the same
assay. Mock, mock-infected samples with no IFN-γ treatment; untreated, BKV-infected
samples with no IFN-γ treatment.
57
250 U/ml IFN-γ resulted in 6.6-fold (P < 0.005) and 12.1-fold (P < 0.004) reductions in
TAg transcript levels, respectively (Figure 2.3B). These findings are consistent with our
previous observations of the protein levels at 96 hours post-infection (Figure 2.2A). The
relative correlation of real time RT-PCR data (measuring the effect of IFN-γ on the
steady-state level of TAg mRNA) and observations made from Western blot analysis
(showing the effect of IFN-γ on the level of viral protein expression) suggest
predominately IFN-γ-mediated effects on transcription. However, IFN-γ may also
mediate an inhibitory effect at the level of translation. Future studies on the mechanism
of inhibition will provide more insight on this subject.
Effect of IFN-γ during infections with different MOIs. To investigate whether the
inhibitory effect of IFN-γ is dependent on levels of input virus, RPTE cells were infected
with five-fold dilutions of BKV. These infections of varying MOIs may represent
different stages of BKV infection in the host. For example, an infection at an MOI of 0.1
might be similar to the subclinical state of persistence seen in immunocompetent hosts,
while an infection at an MOI of 12.5 might be more similar to the reactivation and lytic
infection of BKV preceding the development of PVN in immunosuppressed hosts.
Infected cells were treated with 50 or 250 U/ml IFN-γ and total cell protein was harvested
at two, three, and four days post-infection for analysis of viral protein expression by
Western blotting (Figure 2.4A, 2.4B). Regardless of the MOI used during the infection,
TAg and VP1 expression were strongly inhibited with IFN-γ treatment and peak protein
levels occurred at three days post-infection, similar to the pattern seen in previous
experiments. Not surprisingly, levels of TAg and VP1 expression were higher in samples
from infections with more input virus, but IFN-γ treatment still mediated a strong
58
Figure 2.4. Effect of IFN-γ during infections with different MOIs. RPTE cells were
infected with the TU strain of BKV at MOIs of 12.5, 2.5, 0.5, or 0.1 and treated with 50
U/ml (Figure 2.4A, 2.4B, 2.4C) or 250 U/ml (Figure 2.4C) IFN-γ at three to six hours
post-infection. A and B) Total cell protein was harvested at two, three, and four days
post-infection and analyzed by Western blot, probing for TAg, VP1, and GAPDH. The
analysis of total cell protein harvested from infection at an MOI of 0.5 was repeated in
Figure 2.4B for direct comparison to samples from infection at an MOI of 0.1. M, mockinfected samples with no IFN-γ treatment; dpi, days post-infection. C) Viral lysates were
harvested at T = 0 (after one hour absorption with BKV) and four days post-infection and
viral progeny production was determined by the fluorescent focus assay. Data are
represented as the log of the viral titer in infectious units per ml (IU/ml) and samples
were assayed in triplicate. Untreated, BKV-infected samples with no IFN-γ treatment.
59
reduction in viral gene expression compared to untreated samples with the same infection
conditions. These data suggest that IFN-γ-mediated inhibition may be important for
controlling BKV replication in both immunocompetent and immunosuppressed hosts.
Viral lysates were harvested at four days post-infection from cells infected in
these same conditions and analyzed for viral progeny production using the fluorescent
focus assay. While significant inhibitory effects mediated by IFN-γ were seen in all
treated samples (for MOIs of 12.5, 2.5, 0.5, P < 0.05; for MOI of 0.1, P = 0.05), the effect
was greater when less input virus was used during the infection (Figure 2.4C). Cells
treated with IFN-γ and infected at an MOI of 0.1 showed an almost 80-fold reduction in
viral progeny production, while cells infected at an MOI of 12.5 (7-fold reduction), 2.5
(13-fold reduction), and 0.5 (53-fold reduction) were affected less severely by IFN-γ
treatment. A comparison of viral titers of IFN-γ-treated samples with those of input virus
(T = 0) showed that BKV progeny were produced in the presence of IFN-γ. This
suggests that, despite the inhibitory effects observed, IFN-γ is not driving BKV infection
into a state of latency, in which the complete absence of late gene expression and progeny
production would be expected.
Response of various BKV strains to IFN-γ treatment. To determine whether the
inhibitory effect of IFN-γ on viral replication and gene expression is consistent for
different strains of BKV, we infected RPTE cells with the TU strain, which was used
throughout the above experiments, and two additional strains, Dunlop and Proto-2. Cells
were treated with 50 or 250 U/ml IFN-γ and total cell protein was harvested at two, three,
and four days post-infection for analysis of viral protein expression by Western blotting
(Figure 2.5A). For each strain, TAg and VP1 expression were strongly inhibited by
60
treatment with 50 U/ml IFN-γ and as before, viral protein levels in treated samples
peaked at three days post-infection (Figure 2.5A). Results were similar for samples
treated with 250 U/ml IFN-γ (data not shown). These data suggest that all strains of BKV
may respond similarly to IFN-γ-mediated inhibition of viral gene expression. In addition,
viral lysates were harvested at the initiation of infection (T = 0) and four days postinfection from RPTE cells infected with the three BKV strains and treated with 50 or 250
U/ml IFN-γ (Figure 2.5B). IFN-γ treatment significantly inhibited progeny production
for each of the strains (P < 0.05): viral titer was reduced in the presence of IFN-γ by as
much as 456-fold for infection with the Dunlop strain, 114-fold for infection with the
Proto-2 strain, and 53-fold for infection with the TU strain (Figure 2.5B).
61
Figure 2.5. Response of various BKV strains to IFN-γ treatment. RPTE cells were
infected with the TU, Dunlop, or Proto-2 strains of BKV at an MOI of 0.5 and treated
with 50 U/ml (Figure 2.5A, 2.5B) or 250 U/ml (Figure 2.5B) IFN-γ at three to six hours
post-infection. A) Total cell protein was harvested at two, three, and four days postinfection and analyzed by Western blot, probing for TAg, VP1, and GAPDH. M, mockinfected samples with no IFN-γ treatment; dpi, days post-infection. B) Viral lysates were
harvested at T = 0 (after one hour absorption with BKV) and four days post-infection and
viral progeny production was determined by the fluorescent focus assay. Data are
represented as the log of the viral titer in infectious units per ml (IU/ml) and samples
were assayed in triplicate. Untreated, BKV-infected samples with no IFN-γ treatment.
62
Discussion
Polyomavirus nephropathy is a complication associated with kidney
transplantation resulting from the reactivation of BKV. Our knowledge about the
immune response to BKV is limited, making the management of PVN difficult. The
experiments described in this paper begin to characterize the inhibitory effect of IFN-γ on
the lytic infection of proximal tubule epithelial cells by BKV. We have shown that IFN-γ
specifically inhibits viral gene expression and progeny production in a dose-dependent
manner. In addition, infected cells exposed to IFN-γ have lower levels of TAg transcripts
than untreated cells, suggesting that IFN-γ-mediated inhibition occurs at the level of
transcription. Cells infected with different MOIs of BKV responded similarly to
treatment with IFN-γ; however the more virus present, the weaker the inhibitory effect.
In addition, IFN-γ had a significant inhibitory effect on viral gene expression and
progeny production during infections with three different strains of BKV. These findings
suggest that IFN-γ plays an important role in regulating BKV infection.
It is important to note that samples treated with IFN-γ maintained detectable de
novo viral gene expression and progeny production. It is unclear whether this is a result
of a low level of viral replication in all cells due to incomplete inhibition by IFN-γ, or a
normal level of replication in a fraction of cells that were unresponsive to IFN-γ. In the
latter scenario, IFN-γ may force BKV into a state of latency in responsive cells.
However, the finding that viral replication continued to some extent in the presence of
IFN-γ correlates well with the observation that healthy immunocompetent individuals
shed BKV in their urine periodically throughout their lives. A reduction of IFN-γ
63
production by antigen-specific T lymphocytes in immunosuppressed patients may be, in
part, responsible for reactivation and subsequent development of BKV-associated
disease.
An interesting pattern of viral gene expression was noted in samples exposed to
IFN-γ: protein levels, though severely reduced when compared to untreated samples,
peaked at three days post-infection and then declined until a steady, low level of
expression was reached. We hypothesize that treatment with IFN-γ may either shorten
the duration of TAg expression or force the virus into a persistent or latent state. Since
we observed progeny production at times when TAg expression was either very low or
undetectable by Western blot, a minimal threshold level of TAg seems to be sufficient to
facilitate the progression of a productive BKV infection.
Activation of the IFN-γ signaling cascade as a result of IFN-γ binding to the
surface receptor has a multitude of effects on the cell (Schroder et al., 2004; Sen, 2001).
IFN-γ has overall anti-proliferative and pro-apoptotic effects, mediated primarily by
protein kinase R and IRF-1, respectively. In our studies, we have accounted for these
non-specific cellular effects by normalizing each sample to the levels of a housekeeping
gene, GAPDH or to cell number. IFN-γ signaling also plays a major role in activating
and recruiting cells of the immune system, for example, by up-regulating the surface
expression of MHC class I and II molecules and inducing expression of inflammatory
cytokines. However, the system we used to study the lytic infection of BKV does not
incorporate cells of the immune system and thus eliminates immune activation from the
potential mechanisms of IFN-γ-specific inhibition of BKV replication. In addition, IFN-γ
signaling activates the type I IFN cascade, which establishes a general antiviral
64
environment within the cell. However, the direct treatment of infected cells with IFN-α
has little effect on viral gene expression (Figure 2.1A). The last major category of IFN-γ
signaling effects involves the activation or repression of transcription factors that regulate
the expression of genes. We speculate that the mechanism by which IFN-γ inhibits BKV
replication employs regulatory transcription factors that act on the early and late viral
promoters. However, regulation of the early promoter to inhibit TAg expression alone
would be sufficient to explain a decrease in levels of TAg, VP1, and viral progeny
production, since TAg is a key mediator of the progression of the BKV life cycle. IFN-γactivated transcription factors include members of the STAT and IRF family, ISGF3, cMyc, and c-Jun, as well as other factors affected by downstream elements of the IFN-γ
signaling cascade and subsequently regulated at a later time (de Veer et al., 2001; Der et
al., 1998). We are currently investigating candidate IFN-γ-regulated factors that may be
responsible for the inhibitory effects on BKV.
The non-coding control region (NCCR) is the region of greatest variability
between different strains of BKV. Rearrangements of the transcription factor binding
blocks (designated O-P-Q-R-S in the archetypal BKV strain WW) occur frequently in
tissue culture systems and seem to arise in individuals with high viral loads or
reactivation. We examined the responses of three different strains of BKV to IFN-γ
treatment during lytic infection. BKV TU (NCCR structure: O-P-Q-R1-12-P16-68-Q1-35-R5263-S)
and BKV Dunlop (O-P-P1-7;26-68-P1-64-S) are strains commonly found in
immunocompetent individuals, while BKV Proto-2 (O-P-P1-7;26-68-P-Q1-28-S7-63) was
isolated from the urine of HIV-infected patients (Doerries et al., 1994; Moens and
Rekvig, 2001; Sundsfjord et al., 1994; Sundsfjord et al., 1990). The observation that
65
IFN-γ is able to inhibit viral gene expression and progeny production of three different
strains of BKV with dissimilar NCCR structures suggests that our findings may be
applicable to all strains of BKV. Therefore, we speculate that transcription factor binding
sites common to all three NCCRs (and potentially to most BKV strains) mediate the
inhibitory effect of IFN-γ on the viral promoters.
It is still not understood what factors cause certain kidney transplant patients to
undergo reactivation of BKV and develop PVN, while others on the same
immunosuppressive regimen do not. One possibility involves differences in host
genetics. For example, genetic polymorphisms of the IFNG gene may render some
patients more susceptible to BKV reactivation than others. Polymorphisms are frequent
in the promoter region of the IFNG gene and affect the level of IFN-γ normally produced
in the body (Pravica et al., 1999; Pravica et al., 2000). Recent reports attempt to establish
a correlation between certain IFNG polymorphisms and susceptibility to various diseases,
including HPV-induced cervical cancer (Lai et al., 2005), tuberculosis (Lopez-Maderuelo
et al., 2003), and parvovirus B19 infections (Kerr et al., 2003). It is possible that kidney
transplant patients with IFNG polymorphisms resulting in lower levels of IFN-γ
production are naturally more susceptible to reactivation of BKV under
immunosuppressive therapies. Screening of patients for such polymorphisms may help to
determine the appropriate level of immunosuppression required to prevent graft rejection
but still maintain the ability of the immune system to repress BKV reactivation. In
addition, treatment with IFN-γ may prove beneficial as an alternative to reducing
immunosuppressive therapies upon detection of active BKV replication. There are
precedents for successful interferon therapy: IFN-α has been used to treat a wide-range of
66
cancers, hepatitis B virus, and hepatitis C virus, while IFN-γ has proven effective against
chronic granulomatous disease (Pestka et al., 2004).
In conclusion, IFN-γ has a potent inhibitory effect on BKV gene expression, both
at the level of transcription and translation, and viral progeny production in proximal
tubule cells. This effect is similar for the three different strains of BKV examined and is
more effective at inhibiting gene expression and progeny production in the presence of
less virus. The exact mechanism of IFN-γ-mediated inhibition is not yet known, but we
speculate that the activity of the BKV promoters is regulated by IFN-γ responsive
transcription factors. These findings expand the characterization of the host immune
response to BKV and may lead to new approaches for the prevention of BKV reactivation
in kidney transplant recipients.
67
References
Bohl, D. L., Storch, G. A., Ryschkewitsch, C., Gaudreault-Keener, M., Schnitzler, M. A.,
Major, E. O., Brennan, D. C., 2005. Donor origin of BK virus in renal
transplantation and role of HLA C7 in susceptibility to sustained BK viremia. Am
J Transplant 5, 2213-2221.
Briggs, J. P., Kriz, W., Schnermann, J. B., 2001. Overview of renal function structure.
3rd edition ed. In (A. Greenberg, Ed.),"Primer on Kidney Diseases", Academic
Press, New York, pp. 3-19.
Chen, Y., Trofe, J., Gordon, J., Du Pasquier, R. A., Roy-Chaudhury, P., Kuroda, M. J.,
Woodle, E. S., Khalili, K., Koralnik, I. J., 2006. Interplay of cellular and humoral
immune responses against BK virus in kidney transplant recipients with
polyomavirus nephropathy. J Virol 80, 3495-3505.
Chesters, P. M., Heritage, J., McCance, D. J., 1983. Persistence of DNA sequences of BK
virus and JC virus in normal human tissues and in diseased tissues. J Infect Dis
147, 676-684.
Comoli, P., Azzi, A., Maccario, R., Basso, S., Botti, G., Basile, G., Fontana, I., Labirio,
M., Cometa, A., Poli, F., Perfumo, F., Locatelli, F., Ginevri, F., 2004.
Polyomavirus BK-specific immunity after kidney transplantation. Transplantation
78, 1229-1232.
Daha, M. R., van Kooten, C., 2000. Is the proximal tubular cell a proinflammatory cell?
Nephrol Dial Transplant 15 Suppl 6, 41-43.
de Veer, M. J., Holko, M., Frevel, M., Walker, E., Der, S., Paranjape, J. M., Silverman,
R. H., Williams, B. R., 2001. Functional classification of interferon-stimulated
genes identified using microarrays. J Leukoc Biol 69, 912-920.
Der, S. D., Zhou, A., Williams, B. R., Silverman, R. H., 1998. Identification of genes
differentially regulated by interferon alpha, beta, or gamma using oligonucleotide
arrays. Proc Natl Acad Sci U S A 95, 15623-15628.
Doerries, K., Vogel, E., Gunther, S., Czub, S., 1994. Infection of human polyomaviruses
JC and BK in peripheral blood leukocytes from immunocompetent individuals.
Virology 198, 59-70.
Drachenberg, C. B., Hirsch, H. H., Ramos, E., Papadimitriou, J. C., 2005. Polyomavirus
disease in renal transplantation: review of pathological findings and diagnostic
methods. Hum Pathol 36, 1245-1255.
68
Gardner, S. D., Field, A. M., Coleman, D. V., Hulme, B., 1971. New human papovavirus
(B.K.) isolated from urine after renal transplantation. Lancet 1, 1253-1257.
Hariharan, S., Cohen, E. P., Vasudev, B., Orentas, R., Viscidi, R. P., Kakela, J.,
DuChateau, B., 2005. BK virus-specific antibodies and BKV DNA in renal
transplant recipients with BKV nephritis. Am J Transplant 5, 2719-2724.
Harlow, E., Crawford, L. V., Pim, D. C., Williamson, N. M., 1981. Monoclonal
antibodies specific for simian virus 40 tumor antigens. J Virol 39, 861-869.
Harlow, E., Whyte, P., Franza, B. R., Jr., Schley, C., 1986. Association of adenovirus
early-region 1A proteins with cellular polypeptides. Mol Cell Biol 6, 1579-1589.
Harms, J. S., Splitter, G. A., 1995. Interferon-gamma inhibits transgene expression driven
by SV40 or CMV promoters but augments expression driven by the mammalian
MHC I promoter. Hum Gene Ther 6, 1291-1297.
Heritage, J., Chesters, P. M., McCance, D. J., 1981. The persistence of papovavirus BK
DNA sequences in normal human renal tissue. J Med Virol 8, 143-150.
Hirsch, H. H., 2005. BK virus: opportunity makes a pathogen. Clin Infect Dis 41, 354360.
Hirsch, H. H., Brennan, D. C., Drachenberg, C. B., Ginevri, F., Gordon, J., Limaye, A.
P., Mihatsch, M. J., Nickeleit, V., Ramos, E., Randhawa, P., Shapiro, R., Steiger,
J., Suthanthiran, M., Trofe, J., 2005. Polyomavirus-associated nephropathy in
renal transplantation: interdisciplinary analyses and recommendations.
Transplantation 79, 1277-1286.
Hirsch, H. H., Knowles, W., Dickenmann, M., Passweg, J., Klimkait, T., Mihatsch, M. J.,
Steiger, J., 2002. Prospective study of polyomavirus type BK replication and
nephropathy in renal-transplant recipients. N Engl J Med 347, 488-496.
Hirsch, H. H., Steiger, J., 2003. Polyomavirus BK. Lancet Infect Dis 3, 611-623.
Humes, H. D., Fissell, W. H., Weitzel, W. F., Buffington, D. A., Westover, A. J.,
MacKay, S. M., Gutierrez, J. M., 2002. Metabolic replacement of kidney function
in uremic animals with a bioartificial kidney containing human cells. Am J Kidney
Dis 39, 1078-1087.
Kerr, J. R., McCoy, M., Burke, B., Mattey, D. L., Pravica, V., Hutchinson, I. V., 2003.
Cytokine gene polymorphisms associated with symptomatic parvovirus B19
infection. J Clin Pathol 56, 725-727.
Knowles, W. A., 2001. The epidemiology of BK virus and the occurrence of antigenic
and genomic subtypes. In (K. Khalili, and G. L. Stoner, Eds.),"Human
69
Polyomaviruses: Molecular and Clinical Perspectives", Wiley-Liss Inc., New
York, pp. 527-559.
Lai, H. C., Chang, C. C., Lin, Y. W., Chen, S. F., Yu, M. H., Nieh, S., Chu, T. W., Chu,
T. Y., 2005. Genetic polymorphisms of the interferon-gamma gene in cervical
carcinogenesis. Int J Cancer 113, 712-718.
Livak, K. J., Schmittgen, T. D., 2001. Analysis of relative gene expression data using
real-time quantitative PCR and the 2-ΔΔCT method. Methods 25, 402-408.
Lopez-Maderuelo, D., Arnalich, F., Serantes, R., Gonzalez, A., Codoceo, R., Madero, R.,
Vazquez, J. J., Montiel, C., 2003. Interferon-gamma and interleukin-10
polymorphisms in pulmonary tuberculosis. Am J Respir Crit Care Med 167, 970975.
Low, J., Humes, H. D., Szczypka, M., Imperiale, M., 2004. BKV and SV40 infection of
human kidney tubular epithelial cells in vitro. Virology 323, 182-188.
Meneguzzi, G., Pignatti, P. F., Barbanti-Brodano, G., Milanesi, G., 1978.
Minichromosome from BK virus as a template for transcription in vitro. Proc Natl
Acad Sci U S A 75, 1126-1130.
Moens, U., Rekvig, O. P., 2001. Molecular biology of BK virus and clinical and basic
aspects of BK virus renal infection. In (K. Khalili, and G. L. Stoner,
Eds.),"Human polyomaviruses: molecular and clinical perspectives", Wiley-Liss,
Inc., New York, pp. 359-408.
Nickeleit, V., Singh, H. K., Mihatsch, M. J., 2003. Polyomavirus nephropathy:
morphology, pathophysiology, and clinical management. Curr Opin Nephrol
Hypertens 12, 599-605.
Pestka, S., Krause, C. D., Walter, M. R., 2004. Interferons, interferon-like cytokines, and
their receptors. Immunol Rev 202, 8-32.
Pravica, V., Asderakis, A., Perrey, C., Hajeer, A., Sinnott, P. J., Hutchinson, I. V., 1999.
In vitro production of IFN-gamma correlates with CA repeat polymorphism in the
human IFN-gamma gene. Eur J Immunogenet 26, 1-3.
Pravica, V., Perrey, C., Stevens, A., Lee, J. H., Hutchinson, I. V., 2000. A single
nucleotide polymorphism in the first intron of the human IFN-gamma gene:
absolute correlation with a polymorphic CA microsatellite marker of high IFNgamma production. Hum Immunol 61, 863-866.
Qin, L., Ding, Y., Pahud, D. R., Chang, E., Imperiale, M. J., Bromberg, J. S., 1997.
Promoter attenuation in gene therapy: Interferon-gamma and tumor necrosis
factor-alpha inhibit transgene expression. Hum Gene Ther 8, 2019-2029.
70
Ritter, T., Brandt, C., Prosch, S., Vergopoulos, A., Vogt, K., Kolls, J., Volk, H. D., 2000.
Stimulatory and inhibitory action of cytokines on the regulation of hCMV-IE
promoter activity in human endothelial cells. Cytokine 12, 1163-1170.
Rozen, S., Skaletsky, H. J., 2000. Primer3 on the WWW for general users and for
biologist programmers. In (S. Krawetz, and S. Misener, Eds.),"Bioinformatics
Methods and Protocols: Methods in Molecular Biology", Humana Press, Totowa,
pp. 365-386.
Schroder, K., Hertzog, P. J., Ravasi, T., Hume, D. A., 2004. Interferon-gamma: an
overview of signals, mechanisms and functions. J Leukoc Biol 75, 163-189.
Sen, G. C., 2001. Viruses and interferons. Annu Rev Microbiol 55, 255-281.
Shinohara, T., Matsuda, M., Cheng, S. H., Marshall, J., Fujita, M., Nagashima, K., 1993.
BK virus infection of the human urinary tract. J Med Virol 41, 301-305.
Sundsfjord, A., Flaegstad, T., Flo, R., Spein, A. R., Pedersen, M., Permin, H., Julsrud, J.,
Traavik, T., 1994. BK and JC viruses in human immunodeficiency virus type 1infected persons: prevalence, excretion, viremia, and viral regulatory regions. J
Infect Dis 169, 485-490.
Sundsfjord, A., Johansen, T., Flaegstad, T., Moens, U., Villand, P., Subramani, S.,
Traavik, T., 1990. At least two types of control regions can be found among
naturally occurring BK virus strains. J Virol 64, 3864-3871.
Wheelock, E. F., 1965. Interferon-like virus-inhibitor induced in human leukocytes by
phytohemagglutinin. Science 149, 310-311.
71
CHAPTER III
TRANSFORMING GROWTH FACTOR-BETA-MEDIATED REGULATION OF
BK VIRUS GENE EXPRESSION
BK virus (BKV) is a human polyomavirus that has a well-established role in
complications following transplantation and immunosuppression (Comoli et al., 2006;
Nickeleit and Mihatsch, 2006). In particular, BKV is the causative agent of
polyomavirus nephropathy (PVN) in up to 10% of kidney transplant recipients, resulting
in loss of graft function in 10-80% of those affected (Hirsch et al., 2006). BKV
reactivation is also associated with hemorrhagic cystitis (HC) in bone marrow transplant
recipients, with approximately 50% of patients with active BKV viruria progressing to
HC (Bedi et al., 1995; Pavlakis et al., 2006). Despite extensive investigation, there has
been limited success in identifying antiviral treatments for BKV reactivation and lytic
infection (Josephson et al., 2006; Trofe et al., 2006). Typically, upon diagnosis of BKV
reactivation and PVN, the immunosuppressive regimen of the patient is reduced to allow
the immune system to fight the viral infection. This approach, however, increases the
risk of graft rejection. Thus it is important to investigate the immune response to BKV to
identify immune components that control viral replication.
72
BKV was first isolated from the urine of a kidney transplant patient with ureteral
stenosis in 1971 (Gardner et al., 1971). Primary infection with BKV occurs early in
childhood, with seroconversion occurring by the age of 10 in up to 80% of the human
population (Knowles, 2006). Following the primary infection, BKV undergoes viremic
dissemination and establishes a lifelong persistent infection primarily in the cells of the
kidney and urinary tract (Chesters et al., 1983; Heritage et al., 1981). More specifically,
tubular epithelial cells of the kidney and epithelial cells of the urinary tract are major sites
of BKV persistence and reactivation (Doerries, 2006; Nickeleit and Mihatsch, 2006).
BKV has a non-enveloped, icosahedral virion composed of three proteins, VP1,
VP2, and VP3, which encapsidate a circular double-stranded DNA genome of
approximately 5.2 kb. The genome can be divided into three distinct regions: the early
region, which contains the coding sequences for large tumor antigen (TAg) and small
tumor antigen (tAg); the late region, which contains the coding sequences for VP1, VP2,
VP3, and agnoprotein; and the non-coding control region (NCCR), which contains the
origin of replication and the viral early and late promoters. The NCCR is used to
distinguish one strain of BKV from another because of the propensity of this region to
acquire point mutations and structural rearrangements (Moens and Van Ghelue, 2005).
BKV strains can be divided into two classes: archetypal (pre-rearranged) strains,
presumed to be the infectious and transmissible virus, and rearranged strains, which are
predominately isolated from tissue biopsy samples (Cubitt, 2006). The NCCRs of
archetypal strains are structurally divided into blocks of transcription factor binding sites,
arbitrarily designated O (142 bp, containing the origin of replication, TATA box, and
TAg binding sites), P (68 bp), Q (39 bp), R (63 bp), and S (63 bp) (Markowitz and
73
Dynan, 1988; Moens et al., 1995). Rearranged BKV strains result from the restructuring
of the archetypal NCCR, such that certain blocks (primarily P, Q, and R) are, in whole or
in part, duplicated or deleted (Moens and Van Ghelue, 2005).
It is commonly reported that the predominant NCCR configuration actively shed
in urine is archetypal (Markowitz et al., 1991; Negrini et al., 1991; Sharma et al., 2007;
Sundsfjord et al., 1999; Takasaka et al., 2004), and that NCCR rearrangements seem to
be isolated more frequently from the tissues and sera of patients with high viral loads
(Boldorini et al., 2001; Gosert et al., 2008; Stoner et al., 2002). In addition, changes in
the NCCR structure arise spontaneously in tissue culture and these rearrangements
enhance the ability of the virus to replicate and transform cells (Rubinstein et al., 1991;
Watanabe and Yoshiike, 1985; Watanabe and Yoshiike, 1986). It is possible that changes
in the viral promoter region can also result in altered pathogenesis, such as a heightened
ability to reactivate or disseminate. Furthermore, rearrangements may affect the cell
tropism of BKV, allowing infection of other cell types in addition to kidney and urinary
epithelial cells. The confounding observation, however, is that as of yet there is no
apparent correlation between BKV NCCR structure and clinical outcome (Sharma et al.,
2007).
TGF-β is a secreted cytokine having three isoforms in mammals (TGF-β1, TGFβ2, TGF-β3), all with similar functions involved in the regulation of cell proliferation,
differentiation, and immune suppression (Feng and Derynck, 2005; Li et al., 2006).
TGF-β is produced by many different cell types, including renal epithelial cells, a major
site of BKV reactivation. Interestingly, the expression of TGF-β is enhanced in the
presence of immunosuppressive therapies commonly administered to renal transplant
74
patients (Khanna et al., 1999a; Khanna et al., 1999b; McMorrow et al., 2005; Shihab et
al., 1996). We hypothesized that TGF-β would have an upregulatory effect on BKV gene
expression and replication, correlating with the evident reactivation of BKV in kidney
transplant recipients.
TGF-β is initially expressed as an inactive complex of precursor polypeptides that
undergoes activation by proteolytic cleavage. Upon maturation, TGF-β can bind to TGFβ receptor II dimers, resulting in the recruitment and phosphorylation of TGF-β receptor I
dimers (Shi and Massague, 2003). Once activated, TGF-β receptor I recruits and
phosphorylates Smad2 and Smad3 proteins, the primary components of the TGF-β
signaling cascade (Massague et al., 2005). Phosphorylated Smad2 and Smad3 proteins
can then form a complex with Smad4, resulting in the nuclear translocation of these
proteins. This complex has some weak intrinsic DNA binding activity of its own, but is
more effective in regulating TGF-β-dependent gene expression in conjunction with
cellular transcriptional co-activators that have high DNA binding activity and specificity
(Brown et al., 2007; Massague and Wotton, 2000; Shi et al., 1998).
In this report, we characterize the effect of TGF-β on the lytic infection of BKV
in renal proximal tubule epithelial (RPTE) cells and on the activity of the viral early
region promoter. We demonstrate that the response to TGF-β-mediated regulation is
dependent on the strain of BKV and thus the NCCR structure. We show that
upregulation by TGF-β maps to a short region of the promoter that most likely contains
two distinct transcription factor binding sites. These findings demonstrate transcriptional
regulation of BKV by a cytokine that is found at elevated levels in transplant patients.
75
Materials and Methods
Cell culture and reagents. Primary human renal proximal tubule epithelial cells (RPTE
cells, Cambrex) were maintained in renal epithelial cell basal medium (REBM, Cambrex)
supplemented with human epidermal growth factor, fetal bovine serum (FBS),
hydrocortisone, epinephrine, insulin, triiodothyronine, transferrin, and GA-1000 as
indicated for renal epithelial cell growth medium (REGM, supplements obtained as
REGM SingleQuots, Cambrex). HT-1080 cells (ATCC CCL-121) were maintained in
Dulbecco’s Modified Eagle Medium (Gibco) containing 10% FBS (Cambrex), 100
units/ml penicillin, and 100 µg/ml streptomycin (Cambrex). Both RPTE and HT-1080
cells were grown at 37°C with 5% CO2 in a humidified incubator. Recombinant human
TGF-β1, produced in A293 cells (Peprotech, Inc.), was reconstituted according to
manufacturer’s recommendations and used at a concentration of 10 ng/ml.
Viruses. BKV stocks were prepared from genomic clones of TU (cloned into the EcoRI
site of pGEM-7Zf(-)), and Dunlop and Proto-2 (cloned into the BamHI site of pBR322,
gift of Peter Howley), as previously described (Abend et al., 2007). The resulting crude
viral stocks were titrated by fluorescent focus assay as previously described (Abend et al.,
2007) and the integrity of the NCCR was confirmed by sequencing of PCR products.
Infections. RPTE cells at 70% confluence were infected with the TU, Dunlop, or Proto2 strains of BKV in REGM at an MOI of 0.5 IU/cell (infectious units per cell), incubating
for one hour at 37°C. Viral lysates used for the infection were replaced with fresh
REGM. TGF-β was added at three to four hours post-infection (hpi).
Western blotting. Total cell protein was harvested at three days post-infection (dpi)
using E1A lysis buffer (Harlow et al., 1986) supplemented with 5 μg/ml PMSF, 5 μg/ml
76
aprotinin, 5 μg/ml leupeptin, 0.05 M sodium fluoride, and 0.2 mM sodium orthovanadate.
The Bio-Rad protein assay was used to determine the protein concentration of each lysate
and 10 μg of protein were electrophoresed on an 4-20% Tris-glycine polyacrylamide gel
(Lonza) and analyzed by Western blotting for the expression of viral early protein TAg
and GAPDH as previously described (Abend et al., 2007).
RNA extraction and cDNA synthesis. Total cell RNA was harvested at 24, 36, 48, 72,
and 96 hpi using TRIzol reagent (Invitrogen) according to manufacturer’s instructions.
Samples were treated with DNase I (Promega) to reduce contaminating DNA, and RNA
integrity was confirmed by electrophoresis on an agarose gel. To generate cDNA,
reverse transcription reactions were performed on 1 μg of input RNA using the iScript
cDNA Synthesis Kit (Bio-Rad), according to manufacturer’s instructions.
Real-Time PCR: TaqMan Assay. Primers and probes used to assay TAg and GAPDH
transcript levels are reported previously (Abend et al., 2007). PCR reactions were
performed in a total volume of 25 μl using TaqMan Universal PCR 2x master mix
(Applied Biosystems), 2.5 μl cDNA template, 500 nM of each primer, and 200 nM probe.
The iCycler iQ5 Real-Time Detection System (Bio-Rad) was used for amplification with
the following PCR program: 2 min at 50°C; 10 min at 95°C; 40 cycles of denaturation at
95°C for 15 sec and annealing and extension at 56°C for 1 min. Results are presented as
the fold change in TAg transcript levels, with the relative level observed at 24 hpi,
untreated, arbitrarily set to one. Results are normalized to the levels of GAPDH
transcripts present using the 2-ΔΔC(T) (Livak) method (Livak and Schmittgen, 2001).
Generation of luciferase constructs. The NCCRs of four strains of BKV were
amplified from their genomic clones using the primers
77
Agno1 (5’ AGTGCTAGCGCCTTTGTCCAGTTTAACT 3’) and
LTAg2 (5’ AGTCTCGAGAAATAGTTTTGCTAGGCCTCA 3’), which contain the
restriction endonuclease sites for NheI and XhoI, respectively (underlined). Polymerase
chain reactions utilizing these primers produced 350-400 bp fragments spanning the
NCCR from the start codon of agnoprotein to 35 bp before the TAg start codon. These
fragments were first cloned into the pGEM-T Easy vector (Promega) and then subcloned
into the luciferase vector pGL2-basic (Promega) by means of the NheI and XhoI sites. In
these resulting luciferase constructs (pGL2-TU, pGL2-Dik, pGL2-Dunlop, pGL2-Proto2), the BKV early promoter drives the expression of the firefly luciferase gene.
Site-directed mutagenesis. The following primers were synthesized and HPLC purified
to introduce a point mutation (underlined) at nucleotide 362 (GenBank accession no.
DQ305492) in the BKV TU NCCR: TU-SmtFOR (5’ TCGCAAAACATGT
CTGTGTGGCTGCTTTCCGG 3’), TU-SmtREV (5’ CCGGAAAGCA
GCCACACAGACATGTTTTGCGA 3’). The following primers were synthesized and
HPLC purified to insert 6 bp normally present only in the TU NCCR into the BKV Dik
NCCR (underlined): Dik+6TUFOR (5’ AAACATGTCTGTCTGGCTGC
TTTCCGGTTTCACTCCTTTGG 3’), Dik+6TUREV (5’ CCAAAGGAGTGAAACCG
GAAAGCAGCCAGACAGACATGTTT 3’). Mutagenesis was performed following the
protocol for the QuikChange II Site-Directed Mutagenesis Kit (Stratagene) using the
primer pairs at 1.25 nM each, 100 ng of pGL2-TU or pGL2-Dik as template, 1 mM
dNTPs, and 1.25 U Native Pfu DNA Polymerase (Stratagene) in 25 μl total reaction
volume. The following two-step PCR program was used: 3 min at 95°C; 18 cycles of
denaturation at 95°C for 15 sec and annealing and extension at 68°C for 12 min.
78
Luciferase assays. RPTE or HT-1080 cells were grown in 12-well tissue culture-treated
plates to 60% confluence. Firefly luciferase constructs (pGL2-BKV strain) and a
promoterless control Renilla luciferase plasmid (pRL-Null, Promega) were cotransfected
into RPTE cells at a ratio of 9:1, with a total of 0.6 µg of DNA per well, using the
Effectene Transfection reagent (Qiagen) according to manufacturer’s recommendations.
TGF-β was added at three to four hours post-transfection (hpt) and total cell lysates were
harvested at 48 hpt in 1x Passive Lysis Buffer (Promega). Luciferase assays were
performed on triplicate samples using the Dual-Luciferase Reporter Assay (Promega),
according to manufacturer’s recommendations. Results are expressed as relative light
units (RLU) of firefly luciferase activity normalized to RLU of Renilla luciferase activity.
TGF-β did not significantly affect the levels of Renilla luciferase activity (data not
shown). Statistical significance was determined using a two-tailed Student’s t test
assuming unequal variance, and P values < 0.01 were considered significant.
79
Results and Discussion
TGF-β-mediated regulation of BKV gene expression during infection. We first
investigated the regulatory potential of TGF-β on BKV gene expression during lytic
infection of RPTE cells. We examined the levels of TAg at three dpi in cells infected
with three strains of BKV that differ significantly in their NCCR structures: TU (O-P-QR1-12-P16-68-Q1-35-R52-63-S), Dunlop (O-P-P1-7;26-68-P1-64-S), and Proto-2 (O-P-P1-7;26-68-PQ1-28-S7-63). TGF-β only significantly affected the expression of viral proteins in cells
infected with the TU strain of BKV, as shown by the upregulation of TAg levels (Figure
3.1A). TAg expression remained relatively unchanged in cells infected with the Dunlop
and Proto-2 strains. To demonstrate that the effect was specific to viral gene expression
and not a result of TGF-β-mediated inhibition of epithelial cell proliferation, samples
were also analyzed for levels of GAPDH, a cellular housekeeping gene. These results
indicate that the presence of TGF-β affects BKV gene expression in a strain-specific
manner.
To determine whether TGF-β-mediated regulation occurs during transcription or
translation, we examined the levels of early gene transcripts during the course of BKV
TU infection. Total cell RNA was harvested at 24, 36, 48, 72, and 96 hpi and analyzed
by a real-time RT-PCR assay to detect TAg transcripts in the absence or presence of
TGF-β (Figure 3.1B). Results were normalized to the levels of GAPDH mRNA present
in the samples. TGF-β had a prominent effect on TAg transcription at early time points,
with 2.5-fold upregulation in transcript levels at 24 and 36 hpi, and 3.3-fold upregulation
over untreated cells at 48 hpi. At the later stages of infection the effect of TGF-β was
less pronounced, with only 1.8-fold upregulation in TAg transcripts at both 72 and 96
80
Figure 3.1. TGF-β upregulates BKV TU gene expression during infection. A) RPTE
cells were infected with the indicated strains of BKV at an MOI of 0.5 IU/cell and treated
with 10 ng/ml TGF-β at three to four hpi. Total cell protein was harvested at three dpi
and analyzed by Western blot, probing for TAg and GAPDH. Mock, mock-infected
samples with no TGF-β treatment. B) RPTE cells were infected with the TU strain of
BKV at an MOI of 0.5 IU/cell in the presence or absence of 10 ng/ml TGF-β, and total
cell RNA was prepared at 24, 36, 48, 72, and 96 hpi. Relative TAg transcript levels were
determined using real time RT-PCR, normalizing to levels of GAPDH transcripts in each
sample. One representative experiment is shown; triplicate samples were analyzed in the
same assay. Fold expression of TAg at 24 hpi, untreated was arbitrarily set to one.
Mock, mock-infected samples with no TGF-β treatment.
81
hpi. These results suggest that TGF-β-mediated regulation occurs at the level of
transcription. Furthermore, the timing and limited duration of the upregulation suggest
that transcription factors involved in the TGF-β signaling cascade are responsible for the
effect.
BKV early promoter activity in the presence of TGF-β. We next set out to examine
the effect of TGF-β on the viral early promoter. We cloned the NCCRs of four BKV
strains [TU, Dunlop, Proto-2, Dik (archetypal)] into a luciferase reporter plasmid, such
that the viral early promoter drives the expression of the firefly luciferase gene. This
assay permitted us to examine the effect of TGF-β on promoter activity in the absence of
viral gene expression and other BKV genomic sequences. RPTE cells were cotransfected
with the BKV promoter-driven luciferase plasmids and a promoterless control Renilla
luciferase plasmid in the absence or presence of TGF-β, and lysates were assayed for
luciferase activity at 48 hpt (Figure 3.2A). The results confirmed that only the TU early
promoter was affected by TGF-β treatment, while the other three viral promoters showed
no significant change in activity in the presence of TGF-β (1.4-, 1.1-, and 1.1-fold change
for Dik, Dunlop, and Proto-2 promoters, respectively). The 2.5-fold upregulation (P =
0.009) in TU promoter activity was similar to that seen during BKV infection of RPTE
cells (Figure 3.1). The results of this assay suggest that the effect of TGF-β is mediated
solely at the promoter and is therefore driven by specific transcription factors.
The regulation of BKV early promoter activity by TGF-β signaling may also
depend on the cell type examined. The luciferase assay described above was also
performed in HT-1080 cells, a human fibrosarcoma cell line (Figure 3.2B). The TU and
Dik promoters responded similarly to TGF-β treatment in both cell types: TU promoter
82
Figure 3.2. BKV early promoter activity in the presence of TGF-β. A) RPTE cells
were cotransfected with BKV early promoter-firefly luciferase constructs and a
promoterless control Renilla luciferase plasmid. TGF-β was added at three to four hpt
and total cell lysates were harvested at 48 hpt. Luciferase assays were performed on
triplicate samples and data are represented as relative light units (RLU) of firefly
luciferase activity, normalized to RLU of Renilla luciferase activity. Data shown
represent results obtained from three independent experiments. B) HT-1080 cells were
cotransfected with BKV early promoter-firefly luciferase constructs and a promoterless
control Renilla luciferase plasmid and assayed as described in (A). Data shown represent
results obtained from three independent experiments.
83
activity was upregulated by 3.1-fold (P < 0.001) and Dik activity remained relatively
unchanged (1.6-fold decrease, P < 0.001) in the HT-1080 cells. However, the Dunlop
and Proto-2 promoters were repressed in the presence of TGF-β in HT-1080 cells by 2.6and 3.1-fold respectively (P < 0.001 and P = 0.002, respectively). In addition to cells of
the kidney and urinary tract, BKV sequences have been isolated from peripheral blood
mononuclear cells (Chatterjee et al., 2000; Doerries et al., 1994), tonsils (Goudsmit et al.,
1982), and brain (De Mattei et al., 1995; Elsner and Doerries, 1992). Thus, TGF-βmediated signaling may differentially regulate BKV during infections of other cell types.
Having mapped the effect of TGF-β to the BKV promoter, we wanted to examine
the differences between the TU NCCR and the other three NCCRs used in our studies.
We performed an alignment of the four BKV NCCRs and observed that there was one
region (7 bp, starred nucleotides) of the TU NCCR that did not align to any region of the
Dik, Dunlop, or Proto-2 NCCRs (Figure 3.3). We used the MatInspector transcription
factor binding site prediction program (Cartharius et al., 2005) to analyze the potential
binding sites in the viral promoters. Within this unique sequence in the TU NCCR, the
program predicted a binding site for the transcription factor ZEB-1/AREB6 with a high
probability. There were no predicted ZEB-1 binding sites in the other three NCCRs.
ZEB-1 has been reported to interact with Smad3 and mediate TGF-β-dependent gene
regulation (Postigo, 2003; Postigo et al., 2003). Interestingly, in both the TU and Dik
NCCRs the MatInspector program also predicted an adjacent binding site for the
transcription factor Smad3 with a high probability. The proximity of the predicted
84
Figure 3.3. Alignment of BKV NCCRs. The segments of the NCCRs (TU, Dik,
Dunlop, Proto-2) from the nucleotide before the start codon of agnoprotein through the P
block directly adjacent to the O block are shown. The O block is not shown because it is
highly homologous between strains. Bold type indicates nucleotides that are identical in
at least three of the strains. Underlined regions indicate the predicted Smad3 and ZEB-1
binding sites. Dots indicate nucleotides not found within a sequence. The starred
nucleotides indicate the region of the TU NCCR that does not align anywhere in the Dik,
Dunlop, or Proto-2 NCCRs. The early promoter is read from left to right (top to bottom),
in the direction of the early coding region. The late promoter is read from right to left
(bottom to top), in the direction of the late coding region.
85
ZEB-1 site to the predicted Smad3 site in the TU NCCR suggests that a Smad3-ZEB-1
complex could actively bind to the promoter and regulate early gene transcription.
TU promoter sequences required for TGF-β-mediated regulation. We used sitedirected mutagenesis to modify the luciferase plasmids to define the exact sequences
required for TGF-β-dependent activation of the early promoter (Figure 3.4A). Starting
with the pGL2-TU plasmid as a template, we introduced a single base change (C to G) in
the core of the predicted Smad3 binding site, resulting in the pGL2-TU-Smt plasmid.
This mutation has previously been shown to effectively disrupt the ability of Smad
proteins to bind to the DNA (Jonk et al., 1998). In a similar manner, we modified the
pGL2-Dik plasmid, which contains a predicted Smad3 binding site identical to that found
in the TU NCCR, lacks a predicted ZEB-1 binding site, and is not affected by TGF-β.
Using site-directed mutagenesis, we introduced six nucleotides (GGTTTC) to place a
predicted ZEB-1 binding site in the same relative position and orientation as in the TU
NCCR (pGL2-Dik+6TU).
The wildtype and mutated luciferase reporter plasmids were transfected into
RPTE cells in the absence or presence of TGF-β, and lysates were assayed for luciferase
activity at 48 hpt (Figure 3.4B). Similar to previous results, the TU early promoter was
upregulated by 2.2-fold (P = 0.009) in the presence of TGF-β, while the Dik promoter
showed little change in activity (1.2-fold increase). The mutant TU promoter was also
unresponsive to TGF-β treatment (1.2-fold decrease), suggesting that the putative Smad3
site is required for TGF-β-dependent promoter activation and that the putative ZEB-1 site
alone is not sufficient. In addition, the mutant Dik promoter was upregulated by 2.7-fold
(P = 0.002) in the presence of TGF-β, suggesting that both the predicted Smad3 and
86
Figure 3.4. Regions of the TU promoter required for TGF-β-mediated regulation of
BKV. A) Schematic representation of promoter constructs. TU-Smt has a single base
change in the core of the predicted Smad3 binding site (hatched). Dik+6TU has a 6 bp
insertion that creates the predicted ZEB-1 binding site. The early promoter is read from
left to right, with the start codon for the firefly luciferase gene following the right end of
the promoter. B) RPTE cells were cotransfected with wildtype or mutant BKV early
promoter-firefly luciferase constructs and a promoterless control Renilla luciferase
plasmid. TGF-β was added at three to four hpt and total cell lysates were harvested at 48
hpt. Luciferase assays were performed on triplicate samples and data are represented as
relative light units (RLU) of firefly luciferase activity, normalized to RLU of Renilla
luciferase activity. Data shown represent results obtained from three independent
experiments.
87
ZEB-1 sites are necessary for upregulation. It is important to note that the mutations did
not significantly change the basal activities of the promoters (Figure 3.4B: compare
untreated, TU to TU-Smt; untreated, Dik to Dik+6TU). These results suggest that
regulation depends directly on the presence of binding sites for transcription factors
involved in TGF-β signaling in the viral early promoter, and that the absence of such sites
results in unresponsiveness to TGF-β.
The MatInspector binding site predictions and the results of the luciferase assays
strongly suggested that Smad3 and ZEB-1 regulate the TU promoter. However, our
attempts to show Smad3 and ZEB-1 binding to the viral promoter using three different
assays, electrophoretic mobility shift assay (EMSA), anchored transcriptional promoter
assay (Ravichandran and Major, 2006), and siRNA-mediated knockdown of Smad3 and
ZEB-1, failed to demonstrate conclusively Smad3 and ZEB-1 binding to, or activation of,
the TU promoter in the presence of TGF-β (data not shown). It remains possible that
Smad3 is acting with another transcription factor, or that the assays simply were not well
suited for demonstration of binding of these particular factors in our system. It will be of
interest to define the nature of the factors mediating the TGF-β response in the future.
Previous studies have shown the relevance of Smad-dependent TGF-β signaling
during viral infections. For example, JC virus (JCV)-infected oligodendrocytes have
elevated levels of TGF-β, Smad3, and Smad4, and chloramphenicol acetyl transferase
assays demonstrate the activating effect of Smad protein overexpression on the early and
late viral promoters (Enam et al., 2004). Recently, it was shown that TGF-β stimulates
JCV replication and that MAPK kinase (MEK) inhibitors can block this effect, indicating
a role for activated downstream transcription factors in TGF-β-mediated upregulation
88
(Ravichandran et al., 2007). Epstein-Barr virus (EBV), a ubiquitous human
gammaherpesvirus that, like BKV, has both latent and lytic stages of infection, is also
regulated by TGF-β-mediated signaling. Studies have shown that TGF-β treatment
drives the reactivation of EBV lytic infection from latently-infected B cells (di Renzo et
al., 1994) and epithelial cells (Fukuda et al., 2001). These and other reports have
established the relevance of TGF-β and Smad3 in viral regulation (Li et al., 2006; Reed,
1999).
Our studies of cytokine-mediated regulation of BKV gene expression and
replication are targeted at understanding the process of viral reactivation in
immunosuppressed transplant patients. Previously, it has been shown that common
immunosuppressive therapies used in renal transplantation patients, such as cyclosporine
and tacrolimus, result in upregulation of TGF-β expression in various cell types,
including kidney epithelial cells (Khanna et al., 1999a; Khanna et al., 1999b; McMorrow
et al., 2005; Shihab et al., 1996). Therefore, we hypothesized that TGF-β-mediated
upregulation of the TU strain of BKV would result in an enhanced ability to reactivate in
renal transplant recipients and, consequently, an increase in virulence of this particular
strain. However, there are few reports of the TU strain in clinical isolates (Sundsfjord et
al., 1994; Sundsfjord et al., 1990), and we have been unable to identify other naturally
occurring strains of BKV that contain both the sequences in the TU promoter that are
required for TGF-β-mediated activation. Nevertheless, the predicted Smad3 site occurs
in a number of other strains [MAN10B (GenBank accession no. DQ176633), WWT
(M34048), URO1 (U33549), TC-3 (AF164514), AS (M23122), TW-2 (AB213487),
SJH85B (DQ176634), T2R.1BKreg-3-4 (AF442893), SA090600 (AF356532)], and
89
Smad3 is known to have many different functional binding partners, including AP-1, cMYC, NF-κB, and Sp1 (Brown et al., 2007), all of which also have binding sites in the
NCCR. The results of the luciferase assays using mutated promoters (Figure 3.4) support
the likelihood that there are two transcription factors involved in regulating the response
to TGF-β. Thus, there is potential for TGF-β-mediated activation or repression of other
BKV strains.
Our studies provide evidence for the transcriptional regulation of BKV early gene
expression by as yet unidentified components of the TGF-β signaling cascade. However,
the postulated contribution of TGF-β to BKV reactivation in transplant patients must be
considered in the context of the overall immune status of those individuals. Previously,
we also described the inhibition of BKV replication by IFN-γ, in a strain-independent
manner (Abend et al., 2007). It is therefore reasonable to posit that reactivation of BKV
in immunocompromised patients may require multiple signals, such as the enhancement
of viral gene expression by TGF-β and the absence of IFN-γ-mediated inhibition of
replication, as well as others. Understanding the contributions of each signal to the
outcome of the infection will be critical for building a complete picture of the interaction
between BKV and the host immune system.
90
References
Abend, J. R., Low, J. A., Imperiale, M. J., 2007. Inhibitory effect of gamma interferon on
BK virus gene expression and replication. J Virol 81, 272-279.
Bedi, A., Miller, C. B., Hanson, J. L., Goodman, S., Ambinder, R. F., Charache, P.,
Arthur, R. R., Jones, R. J., 1995. Association of BK virus with failure of
prophylaxis against hemorrhagic cystitis following bone marrow transplantation.
J Clin Oncol 13, 1103-1109.
Boldorini, R., Omodeo-Zorini, E., Suno, A., Benigni, E., Nebuloni, M., Garino, E.,
Fortunato, M., Monga, G., Mazzucco, G., 2001. Molecular characterization and
sequence analysis of polyomavirus strains isolated from needle biopsy specimens
of kidney allograft recipients. Am J Clin Pathol 116, 489-494.
Brown, K. A., Pietenpol, J. A., Moses, H. L., 2007. A tale of two proteins: differential
roles and regulation of Smad2 and Smad3 in TGF-beta signaling. J Cell Biochem
101, 9-33.
Cartharius, K., Frech, K., Grote, K., Klocke, B., Haltmeier, M., Klingenhoff, A., Frisch,
M., Bayerlein, M., Werner, T., 2005. MatInspector and beyond: promoter analysis
based on transcription factor binding sites. Bioinformatics 21, 2933-2942.
Chatterjee, M., Weyandt, T. B., Frisque, R. J., 2000. Identification of archetype and
rearranged forms of BK virus in leukocytes from healthy individuals. J Med Virol
60, 353-362.
Chesters, P. M., Heritage, J., McCance, D. J., 1983. Persistence of DNA sequences of BK
virus and JC virus in normal human tissues and in diseased tissues. J Infect Dis
147, 676-684.
Comoli, P., Binggeli, S., Ginevri, F., Hirsch, H. H., 2006. Polyomavirus-associated
nephropathy: update on BK virus-specific immunity. Transpl Infect Dis 8, 86-94.
Cubitt, C. L., 2006. Molecular genetics of the BK virus. Adv Exp Med Biol 577, 85-95.
De Mattei, M., Martini, F., Corallini, A., Gerosa, M., Scotlandi, K., Carinci, P., BarbantiBrodano, G., Tognon, M., 1995. High incidence of BK virus large-T-antigencoding sequences in normal human tissues and tumors of different histotypes. Int
J Cancer 61, 756-760.
di Renzo, L., Altiok, A., Klein, G., Klein, E., 1994. Endogenous TGF-beta contributes to
the induction of the EBV lytic cycle in two Burkitt lymphoma cell lines. Int J
Cancer 57, 914-919.
91
Doerries, K., 2006. Human polyomavirus JC and BK persistent infection. Adv Exp Med
Biol 577, 102-116.
Doerries, K., Vogel, E., Gunther, S., Czub, S., 1994. Infection of human polyomaviruses
JC and BK in peripheral blood leukocytes from immunocompetent individuals.
Virology 198, 59-70.
Elsner, C., Doerries, K., 1992. Evidence of human polyomavirus BK and JC infection in
normal brain tissue. Virology 191, 72-80.
Enam, S., Sweet, T. M., Amini, S., Khalili, K., Del Valle, L., 2004. Evidence for
involvement of transforming growth factor beta1 signaling pathway in activation
of JC virus in human immunodeficiency virus 1-associated progressive multifocal
leukoencephalopathy. Arch Pathol Lab Med 128, 282-291.
Feng, X. H., Derynck, R., 2005. Specificity and versatility in TGF-beta signaling through
Smads. Annu Rev Cell Dev Biol 21, 659-693.
Fukuda, M., Ikuta, K., Yanagihara, K., Tajima, M., Kuratsune, H., Kurata, T., Sairenji,
T., 2001. Effect of transforming growth factor-beta1 on the cell growth and
Epstein-Barr virus reactivation in EBV-infected epithelial cell lines. Virology 288,
109-118.
Gardner, S. D., Field, A. M., Coleman, D. V., Hulme, B., 1971. New human papovavirus
(B.K.) isolated from urine after renal transplantation. Lancet 1, 1253-1257.
Gosert, R., Rinaldo, C. H., Funk, G. A., Egli, A., Ramos, E., Drachenberg, C. B., Hirsch,
H. H., 2008. Polyomavirus BK with rearranged noncoding control region emerge
in vivo in renal transplant patients and increase viral replication and
cytopathology. J Exp Med 205, 841-852.
Goudsmit, J., Wertheim-van Dillen, P., van Strien, A., van der Noordaa, J., 1982. The
role of BK virus in acute respiratory tract disease and the presence of BKV DNA
in tonsils. J Med Virol 10, 91-99.
Harlow, E., Whyte, P., Franza, B. R., Jr., Schley, C., 1986. Association of adenovirus
early-region 1A proteins with cellular polypeptides. Mol Cell Biol 6, 1579-1589.
Heritage, J., Chesters, P. M., McCance, D. J., 1981. The persistence of papovavirus BK
DNA sequences in normal human renal tissue. J Med Virol 8, 143-150.
Hirsch, H. H., Drachenberg, C. B., Steiger, J., Ramos, E., 2006. Polyomavirus-associated
nephropathy in renal transplantation: critical issues of screening and management.
Adv Exp Med Biol 577, 160-173.
92
Jonk, L. J., Itoh, S., Heldin, C. H., ten Dijke, P., Kruijer, W., 1998. Identification and
functional characterization of a Smad binding element (SBE) in the JunB
promoter that acts as a transforming growth factor-beta, activin, and bone
morphogenetic protein-inducible enhancer. J Biol Chem 273, 21145-21152.
Josephson, M. A., Williams, J. W., Chandraker, A., Randhawa, P. S., 2006.
Polyomavirus-associated nephropathy: update on antiviral strategies. Transpl
Infect Dis 8, 95-101.
Khanna, A., Cairns, V., Hosenpud, J. D., 1999a. Tacrolimus induces increased expression
of transforming growth factor-beta1 in mammalian lymphoid as well as
nonlymphoid cells. Transplantation 67, 614-619.
Khanna, A. K., Cairns, V. R., Becker, C. G., Hosenpud, J. D., 1999b. Transforming
growth factor (TGF)-beta mimics and anti-TGF-beta antibody abrogates the in
vivo effects of cyclosporine: demonstration of a direct role of TGF-beta in
immunosuppression and nephrotoxicity of cyclosporine. Transplantation 67, 882889.
Knowles, W. A., 2006. Discovery and epidemiology of the human polyomaviruses BK
virus (BKV) and JC virus (JCV). Adv Exp Med Biol 577, 19-45.
Li, M. O., Wan, Y. Y., Sanjabi, S., Robertson, A. K., Flavell, R. A., 2006. Transforming
growth factor-beta regulation of immune responses. Annu Rev Immunol 24, 99146.
Livak, K. J., Schmittgen, T. D., 2001. Analysis of relative gene expression data using
real-time quantitative PCR and the 2-ΔΔCT method. Methods 25, 402-408.
Markowitz, R. B., Dynan, W. S., 1988. Binding of cellular proteins to the regulatory
region of BK virus DNA. J Virol 62, 3388-3398.
Markowitz, R. B., Eaton, B. A., Kubik, M. F., Latorra, D., McGregor, J. A., Dynan, W.
S., 1991. BK virus and JC virus shed during pregnancy have predominantly
archetypal regulatory regions. J Virol 65, 4515-4519.
Massague, J., Seoane, J., Wotton, D., 2005. Smad transcription factors. Genes Dev 19,
2783-2810.
Massague, J., Wotton, D., 2000. Transcriptional control by the TGF-beta/Smad signaling
system. EMBO J 19, 1745-1754.
McMorrow, T., Gaffney, M. M., Slattery, C., Campbell, E., Ryan, M. P., 2005.
Cyclosporine A induced epithelial-mesenchymal transition in human renal
proximal tubular epithelial cells. Nephrol Dial Transplant 20, 2215-2225.
93
Moens, U., Johansen, T., Johnsen, J. I., Seternes, O. M., Traavik, T., 1995. Noncoding
control region of naturally occurring BK virus variants: sequence comparison and
functional analysis. Virus Genes 10, 261-275.
Moens, U., Van Ghelue, M., 2005. Polymorphism in the genome of non-passaged human
polyomavirus BK: implications for cell tropism and the pathological role of the
virus. Virology 331, 209-231.
Negrini, M., Sabbioni, S., Arthur, R. R., Castagnoli, A., Barbanti-Brodano, G., 1991.
Prevalence of the archetypal regulatory region and sequence polymorphisms in
nonpassaged BK virus variants. J Virol 65, 5092-5095.
Nickeleit, V., Mihatsch, M. J., 2006. Polyomavirus nephropathy in native kidneys and
renal allografts: an update on an escalating threat. Transpl Int 19, 960-973.
Pavlakis, M., Haririan, A., Klassen, D. K., 2006. BK virus infection after non-renal
transplantation. Adv Exp Med Biol 577, 185-189.
Postigo, A. A., 2003. Opposing functions of ZEB proteins in the regulation of the
TGFbeta/BMP signaling pathway. EMBO J 22, 2443-2452.
Postigo, A. A., Depp, J. L., Taylor, J. J., Kroll, K. L., 2003. Regulation of Smad signaling
through a differential recruitment of coactivators and corepressors by ZEB
proteins. EMBO J 22, 2453-2462.
Ravichandran, V., Jensen, P. N., Major, E. O., 2007. MEK1/2 inhibitors block basal and
transforming growth factor beta1-stimulated JC virus multiplication. J Virol 81,
6412-6418.
Ravichandran, V., Major, E. O., 2006. Viral proteomics: a promising approach for
understanding JC virus tropism. Proteomics 6, 5628-5636.
Reed, S. G., 1999. TGF-beta in infections and infectious diseases. Microbes Infect 1,
1313-1325.
Rubinstein, R., Schoonakker, B. C., Harley, E. H., 1991. Recurring theme of changes in
the transcriptional control region of BK virus during adaptation to cell culture. J
Virol 65, 1600-1604.
Sharma, P. M., Gupta, G., Vats, A., Shapiro, R., Randhawa, P. S., 2007. Polyomavirus
BK non-coding control region rearrangements in health and disease. J Med Virol
79, 1199-1207.
Shi, Y., Massague, J., 2003. Mechanisms of TGF-beta signaling from the cell membrane
to the nucleus. Cell 113, 685-700.
94
Shi, Y., Wang, Y. F., Jayaraman, L., Yang, H., Massague, J., Pavletich, N. P., 1998.
Crystal structure of a Smad MH1 domain bound to DNA: insights on DNA
binding in TGF-beta signaling. Cell 94, 585-594.
Shihab, F. S., Andoh, T. F., Tanner, A. M., Noble, N. A., Border, W. A., Franceschini,
N., Bennett, W. M., 1996. Role of transforming growth factor-beta 1 in
experimental chronic cyclosporine nephropathy. Kidney Int 49, 1141-1151.
Stoner, G. L., Alappan, R., Jobes, D. V., Ryschkewitsch, C. F., Landry, M. L., 2002. BK
virus regulatory region rearrangements in brain and cerebrospinal fluid from a
leukemia patient with tubulointerstitial nephritis and meningoencephalitis. Am J
Kidney Dis 39, 1102-1112.
Sundsfjord, A., Flaegstad, T., Flo, R., Spein, A. R., Pedersen, M., Permin, H., Julsrud, J.,
Traavik, T., 1994. BK and JC viruses in human immunodeficiency virus type 1infected persons: prevalence, excretion, viremia, and viral regulatory regions. J
Infect Dis 169, 485-490.
Sundsfjord, A., Johansen, T., Flaegstad, T., Moens, U., Villand, P., Subramani, S.,
Traavik, T., 1990. At least two types of control regions can be found among
naturally occurring BK virus strains. J Virol 64, 3864-3871.
Sundsfjord, A., Osei, A., Rosenqvist, H., Van Ghelue, M., Silsand, Y., Haga, H. J.,
Rekvig, O. P., Moens, U., 1999. BK and JC viruses in patients with systemic
lupus erythematosus: prevalent and persistent BK viruria, sequence stability of the
viral regulatory regions, and nondetectable viremia. J Infect Dis 180, 1-9.
Takasaka, T., Goya, N., Tokumoto, T., Tanabe, K., Toma, H., Ogawa, Y., Hokama, S.,
Momose, A., Funyu, T., Fujioka, T., Omori, S., Akiyama, H., Chen, Q., Zheng, H.
Y., Ohta, N., Kitamura, T., Yogo, Y., 2004. Subtypes of BK virus prevalent in
Japan and variation in their transcriptional control region. J Gen Virol 85, 28212827.
Trofe, J., Hirsch, H. H., Ramos, E., 2006. Polyomavirus-associated nephropathy: update
of clinical management in kidney transplant patients. Transpl Infect Dis 8, 76-85.
Watanabe, S., Yoshiike, K., 1985. Decreasing the number of 68-base-pair tandem repeats
in the BK virus transcriptional control region reduces plaque size and enhances
transforming capacity. J Virol 55, 823-825.
Watanabe, S., Yoshiike, K., 1986. Evolutionary changes of transcriptional control region
in a minute-plaque viable deletion mutant of BK virus. J Virol 59, 260-266.
95
CHAPTER IV
PRELIMINARY RESULTS ON THE CHARACTERIZATION OF
INTERFERON-GAMMA-MEDIATED REGULATION AND ARCHETYPE
BK VIRUS REPLICATION IN A TISSUE CULTURE SYSTEM
BK virus (BKV) is a ubiquitous pathogen that infects nearly the entire population
by the age of 10 (Knowles, 2001; Knowles et al., 2003). The transmission of BKV is not
yet characterized but is thought to occur by either the respiratory or urino-oral route.
Following a typically subclinical primary infection, BKV is able to disseminate and
establish an infection of kidney epithelial cells, particularly proximal tubule epithelial
cells, and the urothelium (Chesters et al., 1983; Heritage et al., 1981). It is within these
cells that the virus persists throughout the life of the host. Although BKV does not
typically cause disease in healthy individuals, the virus is periodically shed in the urine
(Knowles, 2001; Zhong et al., 2007). In immunosuppressed patients, BKV can cause
severe disease by reactivating to a robust lytic infection. Approximately 10% of renal
transplant recipients undergo BKV reactivation that develops into polyomavirus
nephropathy (PVN), a lytic infection of kidney epithelial cells that can result in loss of
function or destruction of the graft. In addition, approximately 10% of bone marrow
transplant recipients suffer from BKV reactivation, resulting in hemorrhagic cystitis
(HC), a painful infection of bladder epithelial cells characterized by hematuria.
96
BKV is a member of the polyomavirus family and is highly homologous to JC
virus (JCV), the other human polyomavirus, and SV40, the well-studied simian virus
(Cubitt, 2006; Imperiale, 2001). BKV has a small (40-45 nm), nonenveloped icosahedral
virion and a circular, double-stranded DNA genome of approximately 5.2 kb. The
genome can be divided into three major regions: the early coding region, which contains
the genes for large tumor antigen (TAg), small tumor antigen (tAg), and truncated tumor
antigen (truncTAg; D. Das, A. Joseph, J. Abend, D. Campbell-Cecen, and M. Imperiale,
in preparation); the late coding region, which contains the genes for capsid proteins VP1,
VP2, and VP3, and agnoprotein, which does not yet have a defined function; and the noncoding control region (NCCR), which contains the origin of replication and the early and
late promoters, which function in a bidirectional manner. In the nucleus of the infected
cell, BKV genomic DNA becomes associated with the cellular histones H2A, H2B, H3
and H4 to form a viral minichromosome that is then packaged into virions (Meneguzzi et
al., 1978).
This chapter will describe preliminary results from two areas of research that
follow from the work presented in Chapters II and III. The first area of research is a
continuation of the study of interferon-gamma (IFN-γ) mediated inhibition of BKV
replication, with a focus on the characterization of chromatin remodeling events. The
second area of research is aimed at identifying factors that are required for archetype
BKV strains to replicate in a tissue culture system. It is possible that our findings in this
area will also be linked to chromatin remodeling events, implicating chromatin structure
as an important mechanism for regulating BKV replication.
97
Materials and Methods
Cell culture and reagents. Primary human renal proximal tubule epithelial cells (RPTE
cells, Cambrex) were maintained for up to six passages in renal epithelial cell basal
medium (REBM, Cambrex) supplemented with human epidermal growth factor, fetal
bovine serum (FBS), hydrocortisone, epinephrine, insulin, triiodothyronine, transferrin,
and GA-1000 as indicated for renal epithelial cell growth medium (REGM, supplements
obtained as REGM SingleQuots, Cambrex). RPTE cells were grown at 37°C with 5%
CO2 in a humidified incubator. Recombinant human IFN-γ (Peprotech, Inc) was
reconstituted according to manufacturer’s recommendations and used at 50 U/ml or 250
U/ml. Trichostatin A (TSA, Sigma) and sodium butyrate (NaB, Sigma) were
reconstituted according to manufacturer’s instructions and used at 1.32 μM and 5 mM,
respectively.
Viruses and infections. The genome of BKV strain TU was cloned into pGEM-7zf(-) at
the EcoRI site. The genome of BKV strain Dik (archetype) was cloned into pBR322 at
the BamHI site (gift of J. Lednicky). BKV TU viral stocks were prepared and titrated as
previously described (Abend et al., 2007). RPTE cells were infected at 70% confluence
with BKV TU at an MOI of 0.5 IU/cell (infectious units per cell) and incubated at 37°C
for one hour. The viral lysate used for infection was then replaced with fresh REGM.
The dose and time of treatment with IFN-γ, TSA, or NaB is described in the figure
legends.
Transfections. BKV genomic clones were prepared for transfection as follows: pGEM7-TU and pBR322-Dik were digested with EcoRI or BamHI, respectively, recircularized
98
with T4 DNA ligase, and phenol-chloroform extracted. RPTE cells were transfected at
50% confluence using Effectene (Qiagen) and the following conditions: 0.6 μg DNA per
well of a 12-well plate with a DNA to Effectene ratio of 1:20. In the cotransfection
experiment (Figure 4.7), cells were transfected with 0.5 μg of BKV genomic DNA and
0.1 μg of the TAg expression plasmid (pcDNA3.1 with TAg cDNA cloned into the
BamHI and EcoRV sites).
Low molecular weight DNA preparations. Low molecular weight DNA was isolated
from cells using a modified Hirt protocol (Hirt, 1967). Cells were collected and pelleted
at 300 x g for 5 min at 4°C. Cell pellets were resuspended in Lysis buffer (0.6% SDS, 10
mM EDTA) and incubated at room temperature (RT) for 20 min. Sodium chloride
(NaCl) was added to a final concentration of 1.4 M and the samples were incubated
overnight at 4°C to precipitate the cellular genomic DNA, which was then pelleted at
20,000 x g for 20 min at 4°C. The supernatants were incubated with 1.5 μg/μl Pronase E
and 150 ng/μl RNAse A for one hour at 37°C. Samples were phenol-chloroform
extracted three times and precipitated overnight at -80°C with 95% ice cold ethanol and
0.3 M sodium acetate. The low molecular weight DNA was pelleted at 20,000 x g for 10
min at 4°C, washed in 70% RT ethanol, and resuspended in 15 μl of nuclease-free water.
Nuclear low molecular weight DNA was prepared by first isolating nuclei from infected
cells using the Dignam protocol (Dignam et al., 1983). Pelleted nuclei were treated as a
cell pellet and the modified Hirt protocol was performed as described above.
Real time quantitation of viral genomes. Low molecular weight DNA samples
prepared for this analysis were spiked with 500 ng pRL-Null plasmid (Promega) directly
after incubation in Lysis buffer to control for sample loss during the procedure. Primers
99
were designed using Primer3 software (Rozen and Skaletsky, 2000) to amplify 125- and
129-bp fragments of the BKV TU NCCR and β-lactamase gene of pRL-Null (Promega),
respectively. Primers were synthesized by Invitrogen and the sequences are as follows:
TUNCCRFOR (5’ CGCCCCTAAAATTCTCTCTT 3’); TUNCCRREV (5’ ATGTCTG
TCTGGCTGCTTTC 3’); RTAmpFOR (5’ TCGCCGCATACACTATTCTC 3’);
RTAmpREV (5’ GCCGCAGTGTTATCACTCAT 3’). Reactions were performed in a
total volume of 25 μl using Power SYBR Green PCR Master Mix (Applied Biosystems),
1 μl 1:125 diluted sample, and 300 nM of each primer. Amplification was performed in
96-well PCR plates (Bio-Rad) using the iCycler iQ5 Real Time Detection System (BioRad) with the following PCR program: 2 min at 50°C; 10 min at 95°C; 40 cycles of
denaturation at 95°C for 15 sec and annealing and extension at 58°C for 1 min. Results
are presented as the fold change in BKV genome copy numbers, with the levels in the
“36 hpi, IFN-γ, Nuc” sample arbitrarily set to one. Results were normalized to levels of
pRL-null (internal plasmid control) using the Livak method (Livak and Schmittgen,
2001).
Western blotting. Total cell protein was harvested at the indicated times post-infection
using EIA lysis buffer (Harlow et al., 1986) supplemented with 5 μg/ml
phenylmethylsulfonyl fluoride, 5 μg/ml aprotinin, 5 μg/ml leupeptin, 0.05 M sodium
fluoride, and 0.2 mM sodium orthovanadate. Samples were electrophoresed, transferred
to a nitrocellulose membrane, and probed with antibodies as previously described (Abend
et al., 2007). The following primary antibodies were used: pAb416 (Harlow et al., 1981)
for detection of TAg expression, p5G6 (gift of D. Galloway) for detection of VP1
expression, #9171 (Cell Signaling Technology) for detection of phospho-STAT1 levels,
100
and Ab8245 (Abcam) for detection of glyceraldehyde-3-phosphate dehydrogenase
(GAPDH) expression.
Southern blotting. Low molecular weight DNA samples were prepared for
electrophoresis by digestion with EcoRI to linearize the viral genome and DpnI, which
only cuts methylated sequences, to distinguish input DNA from the replicated genome.
Samples were electrophoresed on a 1% agarose gel at 100 V and transferred to a
positively charged nylon membrane (Perkin Elmer) by capillary action overnight in 20x
SSC (3 M NaCl, 0.3 M sodium citrate dihydrate). The probe was prepared by digestion
of pGEM-7-TU with PvuII and isolation of the 3.2 kb fragment corresponding to the
early coding region. The Random Primers DNA Labeling System (Invitrogen) was used
to radiolabel 100 ng of the PvuII fragment with [α-32P]dCTP and 20 ng of probe was
added to the hybridization buffer (5x SSC, 1% SDS, 1x Denhardt’s solution, 100 μg/ml
denatured salmon sperm DNA) for incubation overnight at 68°C. The blot was then
washed extensively and exposed to film.
Site-directed mutagenesis and NCCR swap. The following primers were synthesized
and HPLC purified (Invitrogen) to introduce bases changes (in bold) that insert restriction
enzyme sites (underlined) into the BKV genomic clones: NCCRSpeIFOR (5’ GGGGA
AATCACTAGTCTTTTGCAATTTTTGCAAAAATGG 3’); TAgPmlIFOR
(5’ ACACCACCCCCAAAATAACACGTGCTTAAAAGTGGCTTATAC 3’);
NCCRSacIIFOR (5’ GACAAGGCCAAGATTCCGCGGCTCGCAAAACATGTC 3’).
The reverse primers are exactly the reverse complement of the sequences shown.
Mutagenesis was performed according to the protocol for the Quik Change II SiteDirected Mutagenesis Kit (Stratagene) using primer pairs at 1.25 nM each, 100 ng of
101
pGEM-7-TU or pBR322-Dik as template, 1 mM dNTPs, and 1.25 U Native Pfu DNA
polymerase (Stratagene) in 25 μl total reaction volume. The following two-step PCR
program was used: 3 min at 95°C, 18 cycles of denaturation at 95°C for 15 sec and
annealing and extension at 65.5°C for 16 min and 30 sec. The resulting genomic clones
were digested with SpeI and SacII and both the plasmid and NCCR fragments were
isolated by gel extraction. NCCRs were religated into the opposite genomes and the
DNA was prepared for transfection as described above.
102
Preliminary Results on the Characterization of IFN-γ-Mediated Regulation
IFN-γ is a widely known anti-viral cytokine produced primarily by T cells and
NK cells. It is a hallmark of the T helper 1 (Th1) phenotype of T cells and overall has
pro-inflammatory, anti-proliferative, and pro-apoptotic effects on cells. The IFN-γ
receptor is expressed on most cell types, including the kidney epithelial cells used in our
studies. The signaling cascade is initiated by binding of IFN-γ to its cell surface receptor,
which results in the phosphorylation of Janus kinase 1 (JAK1) and JAK2, which are
associated with the cytoplasmic domains of the receptor. Activated JAK1 and JAK2 then
phosphorylate the receptor, which allows recruitment and binding of signal transducer
and activator of transcription 1 (STAT1). Upon binding the receptor, STAT1 is
phosphorylated, which promotes dimerization and translocation to the nucleus. Activated
STAT1 homodimers act as transcription factors in the nucleus to mediate expression of
IFN-γ responsive genes (reviewed in Pestka et al., 2004; van Boxel-Dezaire and Stark,
2007).
In microarray studies, IFN-γ stimulation has been shown to significantly induce
the expression of over 100 genes at early times post-treatment, including a number of
transcription factors such as STAT1, IRF-1, PML, IRF-9, C/EBP, and TEAD4 (Der et al.,
1998). In addition, IFN-γ signaling promotes the activation of many other factors,
including STAT3, AP-1, USF-1, NF-κB, IRF-1, IRF-8, ATF-2, GATA-1, CREB, and
PU.1 (van Boxel-Dezaire and Stark, 2007). These transcription factors can subsequently
regulate the expression of other proteins in a second wave of IFN-γ-mediated
transcription (van Boxel-Dezaire and Stark, 2007).
103
Transcription factors are DNA binding proteins that regulate transcription mainly
by binding to enhancer regions, recruiting other cellular factors and facilitating the
formation of a stable transcription initiation complex at the promoter of a gene.
Transcription factors can also regulate gene expression by binding and recruiting
chromatin remodeling enzymes, histone acetyltransferases (HATs) and histone
deacetylases (HDACs, reviewed in Bhaumik et al., 2007; Struhl, 2006). The N-terminal
tails of histones contain several lysine residues, which have positive charges that interact
with the negatively charged backbone of DNA, forming a closed, condensed chromatin
structure. HATs transfer acetyl groups (-COCH3) from acetyl-coenzyme A to histone
lysine residues, resulting in the neutralization of charged histone tails and an opening of
the chromatin structure. There are many known HATs, including p300/CBP, P/CAF,
Gcn5, ACTR, SRC-1, and TAF130/250. Hyperacetylated histones are typically
associated with genes that are being actively transcribed. HDACs are enzymes that
remove acetyl groups from the lysine residues, resulting in a stronger interaction between
histones and DNA and subsequently condensing the chromatin. There are four classes of
HDACs, with class I (HDAC1, -2, -3, and -8) being the most widely expressed (Adcock,
2006). Hypoacetylated histones are usually associated with genes that have been
silenced. Histones can also be ubiquitinated, phosphorylated, methylated, and
sumoylated; all of these modifications can affect chromatin structure by altering the
interaction between histones and DNA (Bhaumik et al., 2007).
In Chapter II, we described the inhibitory effect of IFN-γ on BKV replication,
mediated primarily at the level of early gene transcription. The inhibition of TAg
expression results in repression of viral DNA replication, late gene expression, and
104
progeny production due to the integral role of TAg in the viral life cycle. We propose
that there are several mechanisms by which IFN-γ could affect viral replication. First, it
is possible that IFN-γ disrupts viral trafficking during entry or delivery of the viral
genome to the nucleus. More likely, we hypothesize that IFN-γ signaling has a direct
effect on transcription initiation by either 1) activating transcriptional repressors that bind
viral DNA and prevent formation of the transcription initiation complex, 2) activating
transcription factors that recruit HDACs and facilitate the condensation of the viral
genome, or 3) activating transcriptional repressors that prevent expression or activation of
HATs. In this chapter, we will first present data demonstrating that IFN-γ does not affect
delivery of the viral genome to the nucleus, suggesting instead a direct effect on
transcription initiation. In addition, we will show that the effect of IFN-γ is long-lasting
and reversible by simultaneous treatment with HDAC inhibitors. While more studies are
needed, these results suggest that IFN-γ mediates chromatin remodeling events on the
BKV minichromosome.
Results and Discussion
As stated above, an alternative explanation for the observed effects of IFN-γ on
BKV replication (see Chapter II) is that IFN-γ signaling inhibits viral trafficking or
delivery of genomic DNA to the nucleus, the site of viral replication. To rule out this
possibility, RPTE cells were infected with BKV and treated with IFN-γ at 3 hours postinfection (hpi). Low molecular weight DNA was harvested either from whole cells or
from isolated nuclei at 36 hpi, approximately when TAg expression is first detectable but
prior to viral DNA replication, and at 96 hpi, a late stage of infection when viral progeny
105
are being produced and released from cells. Samples were analyzed in a real time PCR
assay to quantify the viral genome copies present (Figure 4.1). At 36 hpi, there was no
difference in viral genome copy number in the nuclei of untreated and IFN-γ treated cells,
indicating that IFN-γ has no effect on viral trafficking or DNA delivery to the nucleus.
As expected, there were more viral genomes present in the whole cell sample than in the
nuclear sample (compare 36 hpi UN, Total and 36 hpi UN, Nuc), indicating that a
fraction of viruses had entered the cell but not yet trafficked to the nucleus. At 96 hpi,
there was a 9.8-fold decrease in the number of viral genomes present in nuclei of IFN-γ
treated cells compared to untreated cells. This result is comparable to our previous
observations of IFN-γ inhibition of TAg transcript levels and protein expression (Chapter
II). Thus, IFN-γ does not appear to affect viral trafficking, but instead mediates an
inhibitory effect on BKV early region transcription, consequently leading to a decrease in
gene expression and viral DNA replication.
Previously, we observed an interesting pattern of viral gene expression in the
presence of IFN-γ: TAg and VP1 levels peaked at 72 and 84 hpi, respectively, after which
expression tapered down to a low level. To examine the extent of this effect, RPTE cells
were infected, treated with IFN-γ at 3 hpi, and total cell lysates were harvested at 5, 7, 9,
11, and 13 days post-infection (dpi). Western blot analysis revealed that this low level of
viral gene expression was maintained throughout the duration of the experiment, even
though the cells were only treated once with IFN-γ (Figure 4.2). We would not expect a
cytokine to be stable for this length of time in the media, however it is possible that the
signaling cascade is being activated by the crosstalk between the type I and type II
interferon pathways (Pestka et al., 2004; van Boxel-Dezaire and Stark, 2007).
106
Figure 4.1. IFN-γ does not affect delivery of BKV DNA to the nucleus. RPTE cells
were infected with BKV TU at an MOI of 0.5 IU/cell and treated with 250 U/ml IFN-γ at
three hpi. Low molecular weight DNA was harvested at 36 or 96 hpi from whole cells
(Total) or isolated nuclei (Nuc). Samples were analyzed by real time PCR to determine
the relative number of BKV genomes in each sample. Results are presented as fold
change in BKV genome copy numbers, normalized to the levels of an internal control
plasmid to account for sample loss during preparation. Samples were analyzed in
triplicate and the fold change in BKV genome copies at 36 hpi in the nuclei of cells
treated with IFN-γ (36 hpi, IFN-γ, Nuc) was arbitrarily set to one. UN, untreated; hpi,
hours post-infection.
107
Figure 4.2. IFN-γ-mediated inhibition of viral gene expression is sustained out to
late stages of infection. RPTE cells were infected with BKV TU at an MOI of 0.5
IU/cell and treated with 50 U/ml IFN-γ at three hpi. Total cell lysates were harvested at
5, 7, 9, 11, and 13 dpi. For each sample, 5 μg of protein were electrophoresed on an 8%
SDS-polyacrylamide gel and analyzed by Western blot, probing for TAg, VP1, and
GAPDH. Mock, mock-infected samples with no IFN-γ treatment; dpi, days postinfection.
108
Alternatively, this result may indicate that histones associated with the viral
minichromosome are being deacetylated, which would reduce viral gene expression and
may require a positive signal for initiation of HAT activity. In this scenario, the viral
genome would appear to be stably repressed in the presence of IFN-γ.
We next postulated that pretreatment of cells with IFN-γ would induce an antiviral state such that viral gene expression would be inhibited from the time of genome
delivery to the nucleus. In this experiment, cells were either untreated, treated with IFN-γ
at 3 hpi, treated at 3 hpi and washed at 24 hpi to remove IFN-γ, pretreated for 24 h with
no treatment after infection, or pretreated for 24 h with IFN-γ added after infection and
washed away at 24 hpi. Total cell lysates were harvested at 5 dpi and analyzed by
Western blot probing for TAg, VP1, phosphorylated STAT1, and GAPDH (Figure 4.3).
It appears that any exposure to IFN-γ is sufficient for repression of TAg expression.
There is a clear difference, however, in the levels of VP1 such that pretreatment appears
inhibit expression more than treatment at 3 hpi. Phosphorylated STAT1 levels were
examined to monitor the activity of the IFN-γ signaling cascade. Interestingly, the
samples in which washes were performed to remove IFN-γ from the cells were very
similar to lysates from unwashed cells, indicating a continued activation of this signaling
cascade at 5 days post-treatment. The persistence of the IFN-γ signaling cascade could
explain the prolonged effect we see on viral gene expression.
To determine if IFN-γ regulates BKV through HDAC activity, cells were infected
and then treated simultaneously at 1 hpi with IFN-γ and either trichostatin A (TSA) or
sodium butyrate (NaB), two broad-spectrum HDAC inhibitors. Total cell protein or low
molecular weight DNA was harvested at 4 dpi and analyzed by Western or Southern blot,
109
Figure 4.3. Pretreatment with IFN-γ results in greater inhibition of gene expression.
RPTE cells were infected with BKV TU at an MOI of 0.5 IU/cell and treated with 250
U/ml IFN-γ in the following ways: treated at 3 hpi, treated at 3 hpi and washed at 24 h
post-treatment, pretreated for 24 h, or pretreated for 24 h, treated again directly after
infection, and washed at 24 h post-treatment. Total cell lysates were harvested at 5 dpi
and 8 μg of protein were electrophoresed on an 8% SDS-polyacrylamide gel and
analyzed by Western blot, probing for TAg, VP1, phosphorylated STAT1, and GAPDH.
Mock, mock-infected samples with no IFN-γ treatment; hpi, hours post-infection; dpi,
days post-infection; IFN-γ 24h pre, pretreated cells for 24 h with IFN-γ; remove 24hpt,
washed cells at 24 h post-treatment.
110
respectively (Figure 4.4). TSA treatment alone resulted in increased TAg expression;
however, in cells treated with both TSA and IFN-γ, the levels of TAg were similar to
those of untreated cells. The same result was seen in the Southern blot analysis for viral
DNA replication, indicating that TSA treatment was able to prevent IFN-γ-mediated
inhibition of BKV. These data suggest that histones associated with the viral genome are
being deacetylated in the presence of IFN-γ. In contrast, NaB had little effect on TAg
expression but seemed to strongly inhibit viral DNA replication both in the absence and
presence of IFN-γ. TSA is a hydroxamic acid compound with potent activity against
class I and class II HDACs, while NaB is a short chain fatty acid with a lower potency
and non-specific inhibitory action against HDACs (Adcock, 2006). Assuming that TSA
is acting more specifically on HDACs, these data suggest that IFN-γ inhibition of BKV is
HDAC-dependent. Alternatively, NaB appears to mediate a decrease in the levels of
modified forms of TAg (various higher molecular weight bands seen in untreated and
TSA-treated lanes, Figure 4.4). Similar modifications have been shown to regulate SV40
TAg functions, and could explain the lack of DNA replication.
So far, these experiments are only suggestive of chromatin remodeling and do not
provide information about specific histone modifications or the location of the modified
histones on the viral genome. Future studies using electrophoretic mobility shift assays
(EMSAs) and chromatin immunoprecipitation (ChIP) will reveal these details. We will
not pursue the HDAC inhibitor studies any further for several reasons. First, HDAC
inhibitors are known to induce cell cycle arrest and apoptosis (Johnstone, 2002; Richon et
al., 2000), and thus non-specific effects may confound results after prolonged exposure.
Second, there are reports that HDACs are required for IFN-γ-mediated
111
Figure 4.4. Treatment with HDAC inhibitors restores BKV gene expression and
replication in the presence of IFN-γ. RPTE cells were infected with BKV TU at an
MOI of 0.5 IU/cell and treated with 250 U/ml IFN-γ and/or 1.32 μM trichostatin A (TSA)
or 5 mM sodium butyrate (NaB) at 1 hpi. Total cell lysates or low molecular weight
DNA were harvested at 4 dpi. To assay for early gene expression, 8 μg of protein were
electrophoresed on an 8% SDS-polyacrylamide gel and analyzed by Western blot,
probing for TAg. To assay for viral DNA replication, low molecular weight DNA
samples were run on a 1% agarose gel and analyzed by Southern blotting as described in
Materials and Methods. HDACI, histone deacetylase inhibitor; WB, Western blot; SB,
Southern blot.
112
signaling events, and that HDAC inhibitors are known to interfere with IFN-γ signaling
by preventing STAT1 phosphorylation, nuclear translocation and gene regulation
(Klampfer et al., 2004; Nusinzon and Horvath, 2005). Although treatment with HDAC
inhibitors has not affected phosphorylation of STAT1 in our studies (data not shown), we
only intended to use this approach as a preliminary screen for HDAC activity in the
presence of IFN-γ.
The large number of genes activated and regulated by IFN-γ signaling makes it
difficult to hypothesize what factors are involved in the inhibition of BKV transcription.
While we postulate that the histone modifications will be located within the NCCR, this
400 to 500 bp region contains many predicted and proven binding sites for transcription
factors that could recruit HDACs or other modifying enzymes. Our immediate plans
involve performing EMSAs with short overlapping probes of the NCCR to map the
regions of DNA that bind additional proteins when incubated with nuclear extracts from
IFN-γ treated cells. Then, targeted sequence analysis and supershift assays with specific
antibodies should allow us to identify the factors involved in the regulation.
Concurrently, we will determine the type and location of histone modifications in the
presence of IFN-γ using ChIP. Immunoprecipitation with antibodies for specific
modified histone residues will indicate the type of modification, while PCR for specific
regions of the genome will map the location of the modified histones. These approaches
will also allow us to examine other modifications of viral DNA-associated histones, such
as methylation, phosphorylation, and ubiquitination.
Chromatin remodeling has been reported as a means to regulate transcription and
replication in a number of viral systems. For example, the latent and lytic states of herpes
113
simplex virus are controlled by remodeling of the viral chromatin (Knipe and Cliffe,
2008). For HIV, histone deacetylation at the LTR is responsible for restriction of Tat
expression and transactivation, and thus establishement of latency (Lusic et al., 2003;
Treand et al., 2006). The acetylation of histones associated with the SV40 genome has
been studied in detail, and appears to play a role in the transition from early to late gene
expression (Balakrishnan and Milavetz, 2006; Balakrishnan and Milavetz, 2007). We are
interested in the ability of cytokine-mediated signaling to induce chromatin remodeling
on the viral genome. We expect that IFN-γ signaling represses BKV gene expression
through HDAC activity, making the viral genome less accessible to transcription
initiation complexes. In our studies, however, viral gene expression was never
completely shut off; there was always a low level of TAg and VP1 expression which
could sustain the infection. This could explain the ability of BKV to maintain a persistent
infection in the host, with minimal gene expression to avoid detection by the immune
response. The reduction in IFN-γ levels by immunosuppression could then allow the
viral minichromosome to be opened and transcribed. It is possible that a positive signal
is required for the activation and recruitment of HATs to the viral genome, which would
fit with the increasingly complex set of risk factors for BKV reactivation in
immunosuppressed patients.
114
Preliminary Results on Archetype BK Virus Replication in a Tissue Culture System
There are two major types of BKV strains, archetype and rearranged,
distinguished from each other by NCCR structure. In archetype strains, the NCCR is
divided into five blocks of DNA sequences arbitrarily designated O, P, Q, R, and S
(Markowitz and Dynan, 1988; Rubinstein et al., 1987; Sundsfjord et al., 1994). The O
block contains the origin of replication and the TAg binding sites, while the other blocks
contain many transcription factor binding sites and were defined by the apparent ability
to move together during NCCR rearrangements. The rearranged strains have NCCR
structures that contain partial or full duplications or deletions of these blocks.
Rearranged BKV strain NCCRs always have O and S blocks and at least one P block, but
the Q and R blocks are frequently deleted and additional P blocks are often present
(Cubitt, 2006; Johnsen et al., 1995; Moens and Rekvig, 2001; Moens and Van Ghelue,
2005).
It was previously shown that the archetype BKV strains are functionally different
from the rearranged strains. Archetype viruses are more efficient at transforming rodent
cells than rearranged viruses, while rearranged strains are far more efficient at replication
in tissue culture (Watanabe and Yoshiike, 1982; Watanabe and Yoshiike, 1986). Several
groups have tried to propagate archetype BKV in tissue culture by infection of cells with
clinical samples or transfection of viral genomic DNA. In each case, the result was either
no viral replication or replication with concurrent NCCR rearrangements, indicating
contamination of the clinical sample with rearranged virus or, in the case of transfected
DNA, the derivation of a rearranged strain from the archetype genome (Rinaldo et al.,
2005; Rubinstein et al., 1991; Sundsfjord et al., 1994; Sundsfjord et al., 1990).
115
Interestingly, the archetype strains are by far the most common BKV strains isolated
from clinical samples, both urine and blood of healthy and immunocompromised
individuals (Gosert et al., 2008; Markowitz et al., 1991; Negrini et al., 1991; Sharma et
al., 2007; Sugimoto et al., 1989; Sundsfjord et al., 1999; Takasaka et al., 2004; ter
Schegget et al., 1985). These observations suggest that the archetype strains are not at all
defective at replication, but that the appropriate conditions for their propagation have not
been attained in tissue culture systems.
Previous studies using in vitro reporter assays have demonstrated that the
archetype early promoter has lower transcriptional activity than rearranged early
promoters (Gosert et al., 2008; Markowitz and Dynan, 1988; Markowitz et al., 1990).
Furthermore, BKV promoters with duplications of only the P block have higher
transcriptional activity (Chakraborty and Das, 1989; Deyerle and Subramani, 1988). This
result could be explained by either the presence of important transcription activator
binding sites within the P block, or by the creation of new transcription factor binding
sites at the junctions of duplicated blocks. The archetype promoter is not inhibitory to
transcription initiation; when short P block fragments from a rearranged NCCR were
inserted into an archetype NCCR, the resulting promoter had enhanced transcriptional
activity suggesting the absence of inhibitory elements (Markowitz et al., 1990).
Although functional for transcription, the lower activity of the archetype early promoter
could result in minimal expression of TAg and thus impede the progression of the viral
life cycle.
In the second part of this chapter, we begin to explore some of the factors that
could be required for archetype BKV propagation. We first propose that the cell types
116
used thus far are simply not supportive of archetype replication and therefore attempt to
productively infect RPTE cells. We also consider the possibility that the archetype
genome is somehow more prone to HDAC activity, resulting in the immediate
condensation of the viral genome upon entry into a cell and complete repression of viral
gene expression. To systematically narrow in on the genomic region that confers the
ability of a BKV strain to replicate, we produced chimeric genomes in which the NCCR
from an archetype strain was replaced with the NCCR from a rearranged strain. The
NCCR swap allowed viral DNA replication of an otherwise archetype genome. Next we
examined the functional relevance of the NCCR swap by cotransfecting the archetype
genome with a TAg expression plasmid, and again observed viral DNA replication.
These results provide information about the conditions in the host that promote archetype
BKV strains to replicate.
Results and Discussion
As mentioned above, propagation of archetype BKV has been attempted in
several common cell lines, including HUVEC-C, a human umbilical vein endothelial cell
line (Rinaldo et al., 2005); Vero, an African green monkey kidney epithelial cell line
(Sundsfjord et al., 1990); and HEK, a human embryonic kidney cell line (Rubinstein et
al., 1991; Sundsfjord et al., 1990). We first wondered if archetype BKV would replicate
in a more relevant cell type, such as the primary RPTE cells characterized by our lab
(Low et al., 2004). These cells are major sites of BKV lytic infection in the host during
reactivation and PVN, and thus it seemed likely they would readily support archetype
replication. RPTE cells were transfected with viral DNA from either a rearranged strain,
BKV TU, or an archetype strain, BKV Dik, and viral lysates were harvested at 10 days
117
post-transfection (dpt). The lysates were then used to infect fresh RPTE cells, which
were fixed at 4 dpi and analyzed for TAg expression by immunofluorescence to assay for
the presence of infectious viral progeny in the transfection lysates (Figure 4.5).
Transfection with the BKV TU genome produced abundant viral progeny, as shown by
the numerous TAg positive cells upon infection. Lysates harvested from cells transfected
with the archetype genome, however, contained no infectious viral progeny, as indicated
by the absence of TAg expressing cells upon infection. The epitope recognized by the
monoclonal antibody used in the immunofluorescence assay (WEQ/SWW; Lindner et al.,
1998; amino acids 91-95 of TAg), is intact in the TAg sequence of archetype virus, thus
these results cannot be explained by an inability of the antibody to recognize infected
cells. It is possible, however, that infectious progeny were produced, but that the level of
TAg expression upon infection was below the limit of detection of the assay. These
results suggest that additional factors are required to drive archetype BKV replication in
RPTE cells.
We next wanted to systematically investigate whether different genomic regions
affected the ability of BKV strains to replicate. Using site-directed mutagenesis, we
inserted unique restriction enzyme sites into archetype and rearranged genomic clones
such that the NCCR was flanked by SpeI and SacII sites, and the early coding region was
flanked by SpeI and PmlI sites (Figure 4.6A). These three sites will allow us to swap the
three major regions of the genome between archetype and rearranged strains.
We began our analysis by swapping the region most likely to affect replication:
the NCCR. The resulting chimeric genomes (Arch/TUN, archetype genome with TU
118
Figure 4.5. Archetype BKV does not productively infect RPTE cells. RPTE cells
were transfected with rearranged (BKV strain TU) or archetype (BKV strain Dik)
genomic DNA and viral lysates were harvested at 10 dpt. Lysates were subjected to three
freeze (-80°C)/thaw (37°C) cycles and used to infect fresh RPTE cells. At 4 dpi, cells
were fixed and assayed by immunofluorescence for TAg expression, as described in
(Abend et al., 2007). Top panels, brightfield. Bottom panels, FITC channel.
119
NCCR; TU/ArchN, TU genome with archetype NCCR) were transfected into RPTE cells
alongside the original archetype and rearranged genomes. Low molecular weight DNA
was harvested at 4 dpt, digested with DpnI, and analyzed by Southern blot to detect
replication of viral DNA (Figure 4.6B). In agreement with previous results, we were
unable to detect replication of the archetype DNA. The chimeric archetype genome,
Arch/TUN, however, replicated robustly in RPTE cells, suggesting that the NCCR is
responsible for the block of archetype BKV replication. Similarly, the BKV TU genome
was efficient at DNA replication, while the sample harvested from transfection of the
chimeric rearranged genome, TU/ArchN, contained no DpnI-resistant DNA, indicating a
lack of viral genome replication. Overall, these results suggest that the NCCR plays an
important role in regulating BKV replication. Future studies will address whether the
replication of the chimeric archetype genome, Arch/TUN, is accompanied by expression
of viral proteins and production of infectious virions. In addition, the remaining
combinations of chimeric genomes will be constructed and analyzed for the ability to
replicate in RPTE cells.
Based on these results, we hypothesized that TAg expression is a limiting factor
for archetype BKV replication. The archetype early promoter has lower activity than
rearranged early promoters and replacement of the archetype NCCR with the higher
activity BKV TU NCCR induced viral DNA replication, suggesting that higher levels of
TAg could rescue replication of archetype strains. To examine the effect of TAg
expression on archetype replication, we cotransfected archetype or BKV TU genomic
clones with a TAg expression plasmid. Low molecular weight DNA was harvested at 4
120
Figure 4.6. Rearranged NCCR can promote archetype BKV DNA replication. A)
Cloning strategy for genomic region swaps. B) RPTE cells were transfected with the
four plasmids from part (A). Low molecular weight DNA was harvested at 4 dpt and
assayed by Southern blot for viral DNA replication, as described in Materials and
Methods. Arch, archetype BKV; Arch/TUN, archetype genome with TU NCCR;
TU/ArchN, TU genome with archetype NCCR; hpi, hours post-infection; dpt, days posttransfection; UN, untreated; Mock, no transfection.
121
and 7 dpt, digested with DpnI, and analyzed by Southern blot to detect replication of
viral DNA (Figure 4.7). BKV TU DNA appears to replicate equally as well in the
absence or presence of the TAg expression plasmid. The archetype genome, however,
only replicates in the presence of TAg overexpression, indicating that elevated levels of
TAg can drive archetype infection. The levels of replicated archetype DNA decreased at
7 dpt, most likely due to the limited duration of TAg expression with transient
transfection. In addition, the diminished intensities of the DpnI digested bands at 7 dpt
were likely a result of cell-mediated degradation of foreign DNA.
These results support the hypothesis that TAg levels are a limiting factor for
archetype replication. In addition to providing insight into the requirements for
productive BKV infection, these data will aide in the development of a tissue culture
system to propagate archetype BKV strains. Future plans include using a cell line stably
transformed with TAg to attempt to propagate and study archetype virus during a
productive infection. Finally, our findings provide support for a model of BKV
reactivation proposed in Chapter V (Figure 5.1), in which rearrangement of the NCCR
precedes reactivation of archetype BKV. The rearranged NCCR is required to provide
higher levels of TAg expression, which promotes the replication of archetype virus.
These and future experiments will help us to better understand the process of archetype
BKV persistence and reactivation in the human host.
122
Figure 4.7. Ectopic expression of TAg can facilitate archetype BKV replication.
RPTE cells were transfected with the rearranged (TU) or archetype (Arch) genomic
clones, or cotransfected with a TAg expression plasmid. Low molecular weight DNA
was harvested at 4 or 7 dpt and assayed by Southern blot for viral DNA replication, as
described in Materials and Methods. Mock, no transfection; dpt, days post-transfection;
C, control plasmid digested with EcoRI as a size marker.
123
References
Abend, J. R., Low, J. A., and Imperiale, M. J. (2007). Inhibitory effect of gamma
interferon on BK virus gene expression and replication. J Virol 81, 272-279.
Adcock, I. M. (2006). Histone deacetylase inhibitors as novel anti-inflammatory agents.
Curr Opin Investig Drugs 7, 966-973.
Balakrishnan, L., and Milavetz, B. (2006). Reorganization of RNA polymerase II on the
SV40 genome occurs coordinately with the early to late transcriptional switch.
Virology 345, 31-43.
Balakrishnan, L., and Milavetz, B. (2007). Histone hyperacetylation in the coding region
of chromatin undergoing transcription in SV40 minichromosomes is a dynamic
process regulated directly by the presence of RNA polymerase II. J Mol Biol 365,
18-30.
Bhaumik, S. R., Smith, E., and Shilatifard, A. (2007). Covalent modifications of histones
during development and disease pathogenesis. Nat Struct Mol Biol 14, 1008-1015.
Chakraborty, T., and Das, G. C. (1989). Identification of HeLa cell nuclear factors that
bind to and activate the early promoter of human polyomavirus BK in vitro. Mol
Cell Biol 9, 3821-3828.
Chesters, P. M., Heritage, J., and McCance, D. J. (1983). Persistence of DNA sequences
of BK virus and JC virus in normal human tissues and in diseased tissues. J Infect
Dis 147, 676-684.
Cubitt, C. L. (2006). Molecular genetics of the BK virus. Adv Exp Med Biol 577, 85-95.
Der, S. D., Zhou, A., Williams, B. R., and Silverman, R. H. (1998). Identification of
genes differentially regulated by interferon alpha, beta, or gamma using
oligonucleotide arrays. Proc Natl Acad Sci U S A 95, 15623-15628.
Deyerle, K. L., and Subramani, S. (1988). Linker scan analysis of the early regulatory
region of human papovavirus BK. J Virol 62, 3378-3387.
Dignam, J. D., Lebovitz, R. M., and Roeder, R. G. (1983). Accurate transcription
initiation by RNA polymerase II in a soluble extract from isolated mammalian
nuclei. Nucleic Acids Res 11, 1475-1489.
Gosert, R., Rinaldo, C. H., Funk, G. A., Egli, A., Ramos, E., Drachenberg, C. B., and
Hirsch, H. H. (2008). Polyomavirus BK with rearranged noncoding control region
124
emerge in vivo in renal transplant patients and increase viral replication and
cytopathology. J Exp Med 205, 841-852.
Harlow, E., Crawford, L. V., Pim, D. C., and Williamson, N. M. (1981). Monoclonal
antibodies specific for simian virus 40 tumor antigens. J Virol 39, 861-869.
Harlow, E., Whyte, P., Franza, B. R., Jr., and Schley, C. (1986). Association of
adenovirus early-region 1A proteins with cellular polypeptides. Mol Cell Biol 6,
1579-1589.
Heritage, J., Chesters, P. M., and McCance, D. J. (1981). The persistence of papovavirus
BK DNA sequences in normal human renal tissue. J Med Virol 8, 143-150.
Hirt, B. (1967). Selective extraction of polyoma DNA from infected mouse cell cultures.
J Mol Biol 26, 365-369.
Imperiale, M. J. (2001). The human polyomaviruses: an overview. In "Human
polyomaviruses: molecular and clinical perspectives" (K. Khalili, and G. L.
Stoner, Eds.). Wiley-Liss, Inc, New York, pp. 53-71.
Johnsen, J. I., Seternes, O. M., Johansen, T., Moens, U., Mantyjarvi, R., and Traavik, T.
(1995). Subpopulations of non-coding control region variants within a cell
culture-passaged stock of BK virus: sequence comparisons and biological
characteristics. J Gen Virol 76, 1571-1581.
Johnstone, R. W. (2002). Histone-deacetylase inhibitors: novel drugs for the treatment of
cancer. Nat Rev Drug Discov 1, 287-299.
Klampfer, L., Huang, J., Swaby, L. A., and Augenlicht, L. (2004). Requirement of
histone deacetylase activity for signaling by STAT1. J Biol Chem 279, 3035830368.
Knipe, D. M., and Cliffe, A. (2008). Chromatin control of herpes simplex virus lytic and
latent infection. Nat Rev Microbiol 6, 211-221.
Knowles, W. A. (2001). The epidemiology of BK virus and the occurrence of antigenic
and genomic subtypes. In "Human Polyomaviruses: Molecular and Clinical
Perspectives" (K. Khalili, and G. L. Stoner, Eds.). Wiley-Liss Inc., New York, pp.
527-559.
Knowles, W. A., Pipkin, P., Andrews, N., Vyse, A., Minor, P., Brown, D. W. G., and
Miller, E. (2003). Population-based study of antibody to the human
polyomaviruses BKV and JCV and the simian polyomavirus SV40. J Med Virol
71, 115-123.
125
Lindner, K., Mole, S. E., Lane, D. P., and Kenny, M. K. (1998). Epitope mapping of
antibodies recognizing the N-terminal domain of simian virus large tumor
antigen. Intervirology 41, 10-16.
Livak, K. J., and Schmittgen, T. D. (2001). Analysis of relative gene expression data
using real-time quantitative PCR and the 2-ΔΔCT method. Methods 25, 402-408.
Low, J., Humes, H. D., Szczypka, M., and Imperiale, M. (2004). BKV and SV40
infection of human kidney tubular epithelial cells in vitro. Virology 323, 182-188.
Lusic, M., Marcello, A., Cereseto, A., and Giacca, M. (2003). Regulation of HIV-1 gene
expression by histone acetylation and factor recruitment at the LTR promoter.
EMBO J 22, 6550-6561.
Markowitz, R. B., and Dynan, W. S. (1988). Binding of cellular proteins to the regulatory
region of BK virus DNA. J Virol 62, 3388-3398.
Markowitz, R. B., Eaton, B. A., Kubik, M. F., Latorra, D., McGregor, J. A., and Dynan,
W. S. (1991). BK virus and JC virus shed during pregnancy have predominantly
archetypal regulatory regions. J Virol 65, 4515-4519.
Markowitz, R. B., Tolbert, S., and Dynan, W. S. (1990). Promoter evolution in BK virus:
Functional elements are created at sequence junctions. J Virol 64, 2411-2415.
Meneguzzi, G., Pignatti, P. F., Barbanti-Brodano, G., and Milanesi, G. (1978).
Minichromosome from BK virus as a template for transcription in vitro. Proc Natl
Acad Sci U S A 75, 1126-1130.
Moens, U., and Rekvig, O. P. (2001). Molecular biology of BK virus and clinical and
basic aspects of BK virus renal infection. In "Human polyomaviruses: molecular
and clinical perspectives" (K. Khalili, and G. L. Stoner, Eds.). Wiley-Liss, Inc.,
New York, pp. 359-408.
Moens, U., and Van Ghelue, M. (2005). Polymorphism in the genome of non-passaged
human polyomavirus BK: implications for cell tropism and the pathological role
of the virus. Virology 331, 209-231.
Negrini, M., Sabbioni, S., Arthur, R. R., Castagnoli, A., and Barbanti-Brodano, G.
(1991). Prevalence of the archetypal regulatory region and sequence
polymorphisms in nonpassaged BK virus variants. J Virol 65, 5092-5095.
Nusinzon, I., and Horvath, C. M. (2005). Histone deacetylases as transcriptional
activators? Role reversal in inducible gene regulation. Sci STKE 2005, re11.
Pestka, S., Krause, C. D., and Walter, M. R. (2004). Interferons, interferon-like
cytokines, and their receptors. Immunol Rev 202, 8-32.
126
Richon, V. M., Sandhoff, T. W., Rifkind, R. A., and Marks, P. A. (2000). Histone
deacetylase inhibitor selectively induces p21WAF1 expression and geneassociated histone acetylation. Proc Natl Acad Sci U S A 97, 10014-10019.
Rinaldo, C. H., Hansen, H., and Traavik, T. (2005). Human endothelial cells allow
passage of an archetypal BK virus (BKV) strain-a tool for cultivation and
functional studies of natural BKV strains. Arch Virol 150, 1449-1458.
Rozen, S., and Skaletsky, H. J. (2000). Primer3 on the WWW for general users and for
biologist programmers. In "Bioinformatics Methods and Protocols: Methods in
Molecular Biology" (S. Krawetz, and S. Misener, Eds.). Humana Press, Totowa,
pp. 365-386.
Rubinstein, R., Pare, N., and Harley, E. H. (1987). Structure and function of the
transcriptional control region of nonpassaged BK virus. J Virol 61, 1747-1750.
Rubinstein, R., Schoonakker, B. C., and Harley, E. H. (1991). Recurring theme of
changes in the transcriptional control region of BK virus during adaptation to cell
culture. J Virol 65, 1600-1604.
Sharma, P. M., Gupta, G., Vats, A., Shapiro, R., and Randhawa, P. S. (2007).
Polyomavirus BK non-coding control region rearrangements in health and
disease. J Med Virol 79, 1199-1207.
Struhl, K. (2006). Histone acetylation and transcriptional regulatory mechanisms. Genes
Dev 12, 599-606.
Sugimoto, C., Hara, K., Taguchi, F., and Yogo, Y. (1989). Growth efficiency of naturally
occurring BK virus variants in vivo and in vitro. J Virol 63, 3195-3199.
Sundsfjord, A., Flaegstad, T., Flo, R., Spein, A. R., Pedersen, M., Permin, H., Julsrud, J.,
and Traavik, T. (1994). BK and JC viruses in human immunodeficiency virus
type 1-infected persons: prevalence, excretion, viremia, and viral regulatory
regions. J Infect Dis 169, 485-490.
Sundsfjord, A., Johansen, T., Flaegstad, T., Moens, U., Villand, P., Subramani, S., and
Traavik, T. (1990). At least two types of control regions can be found among
naturally occurring BK virus strains. J Virol 64, 3864-3871.
Sundsfjord, A., Osei, A., Rosenqvist, H., Van Ghelue, M., Silsand, Y., Haga, H. J.,
Rekvig, O. P., and Moens, U. (1999). BK and JC viruses in patients with systemic
lupus erythematosus: prevalent and persistent BK viruria, sequence stability of the
viral regulatory regions, and nondetectable viremia. J Infect Dis 180, 1-9.
127
Takasaka, T., Goya, N., Tokumoto, T., Tanabe, K., Toma, H., Ogawa, Y., Hokama, S.,
Momose, A., Funyu, T., Fujioka, T., Omori, S., Akiyama, H., Chen, Q., Zheng, H.
Y., Ohta, N., Kitamura, T., and Yogo, Y. (2004). Subtypes of BK virus prevalent
in Japan and variation in their transcriptional control region. J Gen Virol 85,
2821-2827.
ter Schegget, J., Sol, C. J. A., Wouters Baan, E., van der Noordaa, J., and van Ormondt,
H. (1985). Naturally occurring BK virus variants (JL and Dik) with deletions in
the putative early enhancer-promoter sequences. J Virol 53, 302-305.
Treand, C., du Chene, I., Bres, V., Kiernan, R., Benarous, R., Benkirane, M., and
Emiliani, S. (2006). Requirement for SWI/SNF chromatin-remodeling complex in
Tat-mediated activation of the HIV-1 promoter. EMBO J 25, 1690-1699.
van Boxel-Dezaire, A. H. H., and Stark, G. R. (2007). Cell type-specific signaling in
response to interferon-gamma. Curr Top Microbiol Immunol 316, 119-154.
Watanabe, S., and Yoshiike, K. (1982). Change of DNA near the origin of replication
enhances the transforming capacity of human papovavirus BK. J Virol 42, 978985.
Watanabe, S., and Yoshiike, K. (1986). Evolutionary changes of transcriptional control
region in a minute-plaque viable deletion mutant of BK virus. J Virol 59, 260266.
Zhong, S., Zheng, H. Y., and Suzuki, M. (2007). Age-related urinary excretion of BK
polyomavirus by non-immunocompromised individuals. J Clin Microbiol 45,
193-198.
128
CHAPTER V
DISCUSSION
Summary of Results
BK virus (BKV) is a ubiquitous human pathogen, infecting nearly the entire
population early in childhood and persisting throughout the lifetime of the host. In
healthy individuals, BKV infection does not cause disease although viral shedding in the
urine is evident periodically. BKV is widely accepted as the causative agent of
polyomavirus nephropathy (PVN) in renal transplant recipients and late-onset
hemorrhagic cystitis (HC) in bone marrow transplant recipients. The increasing
prevalence of these BKV-associated diseases in immunosuppressed patients is likely a
result of more potent immunosuppressive therapies, which may eliminate components of
the immune system that are necessary to keep the virus in a subclinical state. We
hypothesized that cytokines and cytokine-mediated signaling events are important for
regulating BKV replication, and that the loss of cytokine-producing lymphocytes during
immunosuppression allows BKV lytic infection to occur. This dissertation describes our
investigation of the roles of two cytokines, IFN-γ and TGF-β, in regulating BKV gene
expression and replication.
129
In Chapter II, after screening a panel of cytokines and chemokines, we identified
IFN-γ as having a strong inhibitory effect on BKV early and late gene expression in a
dose-dependent manner. In IFN-γ treated cells, we detected significant reductions in
TAg transcripts at four days post-infection, suggesting that IFN-γ-mediated inhibition
was occurring at the level of transcription. IFN-γ did not appear to change the kinetics of
viral gene expression, but treatment of infected cells resulted in a significant decrease in
viral progeny production. Finally, we demonstrated that the effect of IFN-γ was similar
for three different strains of BKV, suggesting that this cytokine is relevant for the
regulation of all BKV strains.
In Chapter III, we investigated the effect of TGF-β, a cytokine that is stimulated
by certain immunosuppressive therapies, on BKV lytic infection. Viral gene expression,
and specifically the activity of the BKV early promoter, was regulated by TGF-β in a
strain-dependent manner. The TU strain of BKV had enhanced early promoter activity in
the presence of TGF-β, while the Dik, Dunlop, and Proto-2 strains were either unaffected
or had decreased promoter activity, depending on the cell type examined. Using sitedirected mutagenesis, we identified a small segment of the TU promoter that is required
for stimulation of activity in response to TGF-β. While these data suggest that TGF-β
may play a role in BKV reactivation, perhaps more importantly the results demonstrate
that BKV strains can respond differently to cytokine treatment.
We are primarily interested in the process of BKV reactivation in kidney
epithelial cells, leading to the development of PVN in renal transplant recipients. The
primary human renal proximal tubule epithelial (RPTE) cell culture system established
by our lab (Low et al., 2004) is well suited to examine BKV replication in this context.
130
As discussed in Chapter I, we hypothesized based on clinical observations that cytokines
regulate BKV during persistence in healthy individuals, and their differential expression
in transplant patients facilitates reactivation. Our studies of IFN-γ- and TGF-β-mediated
regulation are highly relevant and complementary. The primary targets of
immunosuppressive agents are lymphocytes; most immunosuppression regimens include
calcineurin inhibitors, antiproliferative agents, and/or corticosteriods to block the
replication and activation of T cells and B cells. T cells, especially CD4+ T helper 1 cells
and CD8+ cytotoxic T lymphocytes (CTLs), are major producers of IFN-γ and therefore
levels of IFN-γ are reduced during immunosuppression. Concurrently, TGF-β expression
is enhanced in renal proximal tubular epithelial cells by certain immunosuppressive drugs
(Khanna et al., 1999a; Khanna et al., 1999b; McMorrow et al., 2005; Shihab et al., 1996).
In addition, TGF-β is itself immunosuppressive and the signaling cascades of TGF-β and
IFN-γ are antagonistic (Giannopoulou et al., 2006; Ulloa et al., 1999). Given our results
that IFN-γ inhibits BKV replication and TGF-β enhances replication of certain BKV
strains, we could hypothesize that immunosuppression creates a cytokine environment
that is ideal for BKV reactivation.
Our observation that the TGF-β effect is strain-dependent, however, complicates
this hypothesis. There are no data suggesting that specific rearranged strains are
associated with enhanced pathogenesis, although NCCR rearrangements in general may
be (Gosert et al., 2008). We were able to map the TGF-β-responsive elements in the TU
strain; although we were unable to find any other strain with a predicted ZEB-1 binding
site, a number of strains contain putative Smad3 binding sites (listed in Chapater III).
Therefore, it is possible that additional TGF-β-responsive elements exist in other strains,
131
as Smad3 has many documented binding partners (Brown et al., 2007; Feng and Derynck,
2005). Furthermore, the effects of TGF-β are known to be different for epithelial cells,
fibroblasts, and immune cells (Li et al., 2006; Rahimi and Leof, 2007); thus the use of
these cell types in our assays may reveal other TGF-β regulated strains of BKV. In
particular, likely candidates for such experiments include lymphocytes, which may play a
role in the dissemination of BKV during primary infection (Doerries et al., 1994), and
bladder epithelial cells, which are the sites of viral lytic replication during HC.
Finally, it is possible that the importance of TGF-β for BKV reactivation stems
from the immunosuppressive and anti-inflammatory effects of this cytokine.
Specifically, TGF-β inhibits the proliferation and differentiation of naïve T cells into
effector cells, CTLs and T helper cells (Gorelik and Flavell, 2002; Li et al., 2006),
thereby inhibiting expression of IFN-γ. In this scenario, the relevance of TGF-β
signaling in BKV reactivation is indirect: a block in the production of IFN-γ would
alleviate the repression of viral replication, but there is no direct regulation of promoter
activity by TGF-β signaling components. In addition, TGF-β, in conjunction with IL-6,
stimulates the development of T helper 17 cells, named for their ability to produce
members of the IL-17 family of proinflammatory cytokines (Bettelli et al., 2008;
Steinman, 2007; Tato and O'Shea, 2006). In particular, IL-17 has been implicated in
immune-mediated tissue injury. Given that immunosuppression alone is not sufficient to
cause PVN, and instead it appears that some renal tissue damage may also be required,
IL-17 expression and signaling events may play a role in BKV reactivation and disease.
The best way to investigate hypotheses involving the regulation of immune cells
and their effects is to use a small animal model for BKV persistence and reactivation.
132
Our lab has made several attempts to develop such a model. First, we tried to establish a
mouse model of PVN using K virus, a murine polyomavirus that is genetically more
similar to BKV than mouse polyomavirus (Py), most notably because it does not encode
a middle T antigen (Imperiale and Major, 2007). Although previous reports describe the
establishment of a persistent K virus infection in kidney epithelial cells of mice, we were
unable to reproduce these results. Next, we collaborated with a lab interested in
identifying the block to BKV replication in rodent cells, which may allow the subsequent
development of a transgenic mouse that supports BKV lytic replication. This work has
not yet revealed a cellular factor that confers permissiveness of rodent cells to productive
BKV infection. Recently, a mouse model for PVN using Py was reported: mice
underwent renal transplantation followed by infection with Py, and the characteristics and
effects of viral replication were examined (Han Lee et al., 2006). Although this model
demonstrated preferential Py replication in the graft resulting in accelerated graft failure,
it did not incorporate a persistent infection prior to transplantation. In addition, the
significant genetic differences between Py and BKV limit the usefulness of this system,
as only questions about the host response to viral reactivation could be addressed.
Establishment of an accurate small animal model for PVN would be extremely valuable
for the study of BKV replication and the immune system during reactivation.
In Chapter IV, we described the preliminary results from two current areas of
research. First, we have continued to investigate the regulation of BKV by IFN-γmediated signaling events, as this cytokine has inhibitory effects on all viral strains
examined and thus is relevant for an overall understanding of the immune response to
BKV. We demonstrate that IFN-γ does not affect viral trafficking or delivery of the
133
genome to the nucleus, which further supports regulation at the level of transcription.
The inhibition appears to be long-lasting, effective to at least nine days post-infection,
and is stronger if cells are treated with IFN-γ prior to infection. Pretreatment may cause
the cells to adopt an antiviral state in which the factors that regulate BKV are activated
and recruited to the nucleus even before infection; thus, when viral DNA enters the
nucleus it is immediately repressed. We hypothesized that IFN-γ signaling results in a
stable chromatin remodeling event that drives the viral genome into a closed
conformation. Treatment of infected cells with broad spectrum histone deacetylase
(HDAC) inhibitors restores viral gene expression and replication in the presence of IFNγ, supporting this hypothesis.
The second area of research described in Chapter IV is aimed at identifying the
factors required for archetype BKV strains to replicate in tissue culture. While we were
optimistic that RPTE cells, the model system developed by our lab to mimic BKV lytic
infection in the kidney, would support archetype replication, there was no indication of
infectious progeny production or viral DNA replication in our assays. Instead, we began
a systematic approach to identify the region of the archetype genome that is responsible
for viral inactivity. By inserting unique restriction enzyme sites flanking the NCCR and
the early coding region, we can exchange genomic fragments between the rearranged TU
strain and the archetype Dik strain. Substitution of the rearranged NCCR into the
archetype genomic clone promoted robust viral DNA replication. Furthermore,
cotransfection of the archetype genome with a TAg expression plasmid also allowed
replication of viral DNA. These findings suggest that the promoter activity of the NCCR
determines the ability of a viral strain to replicate in tissue culture. We further
134
hypothesized that the NCCRs of rearranged and archetype strains may differ in chromatin
structure, resulting in a difference in promoter activity. Thus, the preliminary results
from both areas of research in Chapter IV will lead to a deeper investigation of chromatin
remodeling events in the context of viral infection.
Major Questions in the Study of BKV Reactivation
There are at least four critical questions that must be addressed to further our
understanding of BKV reactivation and associated disease in the context of
immunosuppression:
1) What factors or conditions allow archetype BKV to replicate in the host but not in
a tissue culture system?
2) What is the pathological relevance of the archetype and rearranged BKV strains?
3) Why are the archetype strains preferentially shed in the urine, while the
rearranged strains are preferentially found in the blood?
4) Ultimately, what are the factors that drive BKV reactivation in kidney transplant
patients?
We will discuss these questions in greater detail in the following paragraphs.
What factors or conditions allow archetype BKV to replicate in the host but not in a
tissue culture system? As described in Chapters I and IV, there have been many studies
demonstrating that archetype BKV strains do not replicate in tissue culture systems
(Rinaldo et al., 2005; Rubinstein et al., 1991; Sundsfjord et al., 1994; Sundsfjord et al.,
1990; Watanabe and Yoshiike, 1982; Watanabe and Yoshiike, 1986). In the human host,
135
however, these viruses replicate quite efficiently, since the vast majority of clinical
isolates, both from healthy individuals and immunosuppressed patients, have archetype
NCCR structures (Gosert et al., 2008; Markowitz et al., 1991; Negrini et al., 1991;
Sharma et al., 2007; Sugimoto et al., 1989; Sundsfjord et al., 1999; Takasaka et al., 2004;
ter Schegget et al., 1985). Based on these observations, it appears that archetype BKV
strains are in no way defective at replication, and instead we can propose that there are
either factors missing that may direct propagation or inhibitory factors present that are
preventing it.
The simplest explanation for these observations is that archetype BKV is only
able to productively infect very specific cell types in the host, and that these cells have
not yet been used for propagation studies. Other viruses have demonstrated such
specificity, notably human papillomaviruses, which replicate in stratified squamous
epithelium in the host and rely on differentiation of the cells for progression through the
viral life cycle (Howley and Lowy, 2007). It was not until the development of complex
raft cultures that researchers were finally able to study and propagate papillomaviruses in
a cell culture system (Asselineau and Pruniera, 1984; McCance et al., 1988). The
restriction of archetype BKV replication is not related to entry or trafficking, since
transfection of viral DNA does not allow propagation (see Chapter IV; Rinaldo et al.,
2005; Rubinstein et al., 1991). Instead, this restriction is likely the result of differential
expression of transcription factors that regulate viral early gene expression. Host cell
restriction by transcription factor expression has been suggested for JC virus (JCV), for
which productive infection of rearranged strains appears to be limited to cells that express
high levels of NF-1/X (Messam et al., 2003; Monaco et al., 2001). Similarly, expression
136
of a specific transcription factor may also regulate archetype BKV replication, and the
permissive cell types have simply not yet been examined.
One promising candidate cell type for susceptibility to archetype strains is
primary bladder epithelial cells. BKV sequences have been detected in normal and
neoplastic bladder epithelium (Monini et al., 1995). Furthermore, the association of BKV
with HC, a bladder infection, and the prevalence of virus in urine samples from healthy
and immunosuppressed individuals suggest susceptibility of these cells to infection. We
are currently investigating whether primary bladder epithelial cells will support
productive archetype BKV infection. However, the bladder and ureter are lined with
transitional epithelium, layers of epithelial cells not unlike stratified squamous
epithelium, which allow the bladder to contract and expand. It is possible that archetype
BKV will require a more complicated cell culture system, like that of human
papillomavirus, to efficiently replicate.
Alternatively, it is possible and perhaps more likely that archetype strains require
a positive factor to initiate replication, one that persistently infected cells are exposed to
only during certain conditions. For example, it has been shown that the S block of the
NCCR contains steroid hormone response elements and treatment of BKV-infected cells
with glucocorticoids, estrogen, or progesterone results in higher viral yields (Moens et al.,
1994). A hormone-dependent enhancement of viral replication would be relevant for
reactivation in kidney and bone marrow transplant patients, who are often treated with
corticosteroids as a part of their immunosuppressive regimen. There is some evidence
suggesting that corticosteroid therapies are associated with an increased risk for BKV
reactivation and PVN (Hirsch et al., 2002; Trofe et al., 2003). In addition, progesterone-
137
mediated enhancement of viral replication may be relevant to regulation of BKV during
pregnancy. Viral shedding is more prevalent in pregnant women compared to the general
population, particularly in the second and third trimesters when the levels of progesterone
peak (Bendiksen et al., 2000; Boldorini et al., 2008; Coleman et al., 1980; Markowitz et
al., 1991). Thus, steriod hormones may be positive factors that stimulate cells to be
permissive to archetype BKV replication.
Another possible factor required for propagation of archetype strains may be
elevated levels of TAg expression. As previously discussed (Chapter I and Chapter IV),
archetype strains are known to have lower early promoter activity than rearranged strains.
It is possible that these viruses do not produce enough TAg to facilitate viral DNA
replication. JCV archetype and rearranged strains have characteristics very similar to
those described for BKV (reviewed in Yogo and Sugimoto, 2001). Propagation of JCV
archetype strains has been demonstrated in cells ectopically expressing JCV TAg (Sock
et al., 1996) and COS-7 cells, which stably express SV40 TAg (Hara et al., 1998),
without the induction of NCCR rearrangements. Based on our preliminary results in
Chapter IV, it is likely that the same will be true for archetype BKV, however, further
studies are needed to demonstrate progeny production and stability of the NCCR.
As proposed in Chapter IV, there may be differences in viral chromatin structure
that could explain the ability of rearranged strains to replicate more efficiently than
archetype strains. For example, the archetype NCCR may contain more or higher affinity
binding sites for transcription factors that recruit HDACs, leading to condensation of the
genome and restriction of transcription and replication. Conversely, rearranged NCCRs
may have additional binding sites that aide in the recruitment of histone
138
acetyltransferases (HATs), leading to an open chromatin structure that promotes
transcription and replication. In either case, the positive factor required for archetype
replication could be one that inhibits HDAC (or stimulates HAT) activity or recruitment.
Many signaling cascades, including those activated in response to cytokines, could be
responsible for such an effect by activating or inducing the appropriate transcription
factors. However, treating cells with HDAC inhibitors may allow efficient propagation
of archetype virus, and the subsequent study of BKV minichromosome structures and
factors that participate in regulation under different conditions.
What is the pathological relevance of the archetype and rearranged BKV strains?
For JCV, the pathological relevance of NCCR structure is quite clear. JCV is the
causative agent of progressive multifocal leukoencephalopathy (PML), a fatal
demyelinating disease that affects severely immunosuppressed individuals (reviewed in
Khalili et al., 2006). Viral genomes containing promoter regions with a conserved linear
structure, designated archetype, are primarily isolated from urine samples, kidney, and
tonsillar tissues of healthy individuals (Agostini et al., 1996; Jeong et al., 2004; Kato et
al., 2004; Tominaga et al., 1992; Yogo et al., 1990). Viral genomes with NCCR
rearrangements, deletions, or amplifications are isolated routinely from the brains of
PML patients and are designated PML-type strains (Loeber and Doerries, 1988; Martin et
al., 1985). Furthermore, PML-type strains only replicate efficiently in human glial cells,
the site of lytic infection and disease, and stromal cells and B lymphocytes, which likely
aide in viral dissemination to the brain (Imperiale and Major, 2007). Thus, NCCR
139
rearrangements are highly associated with JCV lytic infection in the brain and
progression to PML.
Based on the similarities between BKV and JCV, we would predict that NCCR
rearrangements in BKV NCCRs would also correlate with disease progression.
However, the fact that BKV both causes disease and maintains persistence at the same
sites in the host, the kidney and bladder epithelium, creates difficulty in distinguishing
strains associated with disease from those associated with persistence and transmission.
In a recent report, Gosert et al. (2008) made several important observations. First,
rearranged strains were more frequently detected in the plasma than in the urine, although
archetype strains still predominated in both. Second, patients with rearranged BKV had
significantly higher viral loads in the blood. Finally, among patients that had progressed
to PVN, those with rearranged BKV had more inflammation and more extensive cell
damage than patients with archetype BKV, as determined by examination of kidney
biopsies. This is the first study to correlate BKV NCCR rearrangement with more severe
disease in kidney transplant patients.
While it is possible that certain conditions of the host could predispose archetype
BKV to acquire NCCR rearrangements, it seems likely that the changes occur randomly
and sporadically, especially since rearrangements have been inconsistently observed in
parallel cultures in vitro (Rinaldo et al., 2005). The functional relevance of the
rearrangement might be very general: the resulting viral promoter is more active for early
region transcription, resulting in more DNA replication and progeny production. The
higher viral loads then lead to greater cell damage and more severe disease. In
immunocompetent individuals, robust expression of viral antigens could lead to early
140
detection by the immune response and elimination of virus-infected cells before progeny
production. Thus, in healthy individuals, rearranged strains may be at a disadvantage due
to activation of the immune response, while archetype strains may be more adept at
immune avoidance, and thus persistence and replication.
In contrast, it is possible that NCCR rearrangements change the cell tropism of the
virus. This would explain the difficulty in propagating the archetype strains in the same
cells that support replication of rearranged strains. Alternatively, NCCR rearrangements
might change the way the virus is regulated in response to cytokine-mediated signaling or
other factors. As suggested in Chapter III, the presence or absence of transcription factor
binding sites can result in differential regulation of viral promoters in the presence of
cytokines. There is one report suggesting that BKV strains containing a mutated Sp1
binding site are associated with development of hemorrhagic cystitis in bone marrow
transplant recipients (Priftakis et al., 2001). Gosert et al. (2008), however, did not report
any NCCR rearrangements or binding sites that were common among the isolates from
the blood of PVN patients. While differential regulation by cytokines is an appealing
idea, it is not yet supported by clinical data.
It is important to restate that the majority of BKV sequences in clinical samples
have archetype NCCR structures. Gosert et al. (2008) report that 85 and 92% of
sequences in the blood and urine, respectively, were archetype BKV and that
approximately 50% of the patients that developed PVN were infected with archetype
strains. These data argue against a clinical relevance for NCCR rearrangements, as it is
clear that the replication of archetype strains can result in disease. Instead, perhaps
rearranged NCCRs facilitate the replication of archetype strains by providing higher
141
levels of TAg expression in trans (as discussed above). If NCCR rearrangements are
random events, it should be possible to have genomes with changes in the coding regions
as well. Thus, in addition to the functional NCCR rearranged strains, there could also be
non-functional genomes with coding region rearrangements that disrupt viral gene
expression and are defective for replication. One could then envision a population of
viral genomes with highly active rearranged NCCRs and disruptions to the late coding
region. These genomes would express high levels of TAg but no late proteins, and
therefore would be defective for replication themselves; however, the expression of TAg
from these ‘helper’ viruses could facilitate archetype BKV replication. In this scenario,
rearrangements in the NCCR may be indirectly essential for archetype BKV persistence
and reactivation.
Why are the archetype strains preferentially shed in the urine, while the rearranged
strains are preferentially found in the blood? This question is perhaps a bit
misleading: archetype strains of BKV are actually the most common strains found in both
blood and urine samples, as described above and in Chapter I. Interestingly, Gosert et al.
(2008) observed that the viral load of archetype BKV in the urine was significantly
higher than that of rearranged BKV, while the opposite was true in the blood. We can
propose two different explanations for this observation: either the two types of virus are
replicating in different cells or the progeny are being released in different ways.
Based on the inability to propagate archetype strains in cells that are highly
susceptible to infection with rearranged BKV, it is possible that the two types of virus
exhibit different cell tropisms (this idea has been discussed in the sections above). It is
142
necessary to assume, however, that archetype BKV can readily infect kidney epithelial
cells, although it may require specific conditions to productively replicate. Archetype
strains are thought to be the transmitted form of BKV, and the derivation of rearranged
strains from archetype genomes has been demonstrated. Therefore, archetype virus must
be capable of establishing persistence in the kidney. If archetype strains are also able to
infect additional cell types, such as primary bladder epithelial cells, then simultaneous
infection of kidney and bladder epithelial cells could allow archetype progeny to be
released into the blood and urine. If rearranged strains can only infect the kidney
epithelium, then viral progeny may be channeled primarily into the blood.
The alternative explanation is that there is a difference in the release of archetype
and rearranged progeny from the cell. Very little is known about how polyomaviruses
exit the infected cell, but there are reports describing both lytic bursts and viral egress
from intact cells (Imperiale and Major, 2007). Although there has been no direct
comparison made between the replication kinetics of archetype and rearranged virus in
the host, the behavior of these two viruses in tissue culture suggests that archetype BKV
has a slow replication cycle and rearranged strains replicate quickly and robustly.
Perhaps this rapid production of rearranged BKV progeny overwhelms the normal viral
egress pathway and instead virions are released by a lytic burst. The destruction of
epithelial cells may help the virus spread and infect surrounding cells, including the
endothelial cells of nearby blood vessels. Replication in and subsequent lysis of these
cells could release the virus into the bloodstream. In contrast, the slower replication of
archetype strains could promote viral egress, leaving infected cells intact and perhaps
favoring release of virus in the urine instead of the bloodstream. The presence of
143
archetype virus in the blood of immunocompromised patients with PVN may indicate
that extensive and unregulated replication is also possible with archetype strains and will
result in cell lysis.
Differences in viral release could also be mediated by mechanisms of egress.
Epithelial cells are polarized and thus have apical surfaces, which face the lumen, and
basolateral surfaces, which contact the surrounding epithelial cells. Many viruses have
been shown to release progeny preferentially from one of these surfaces, including
hepatitis A (Blank et al., 2000), measles virus (Blau and Compans, 1995), SV40 (Clayson
et al., 1989), and Epstein-Barr virus, for which preferential basolateral release has been
suggested to favor viral dissemination (Chodosh et al., 2000). Rearranged strains may
somehow be defective for viral egress at the apical surface but efficient at release from
the basolateral surfaces, leading to dissemination into the blood. In contrast, archetype
strains may either mediate apical release preferentially from bladder epithelial cells and
basolateral release from kidney epithelial cells, or perhaps have no preference for release.
NCCR rearrangements could affect viral release by regulating replication kinetics, as
described above. Alternatively, rearrangements could result in differential regulation of
agnoprotein expression; very little known about this protein, but it is suggested to be
involved in virus maturation and release (Rinaldo et al., 1998).
Ultimately, what are the factors that drive BKV reactivation in kidney transplant
patients? Currently, the list of potential risk factors for PVN includes older age, male
gender, seropositivity of the donor, seronegativity of the recipient, specific
immunosuppressive drugs, HLA mismatches, lack of the HLA-C7 allele in the donor or
144
recipient, acute rejection episodes prior to development of PVN, and low numbers of
BKV-specific IFN-γ-producing T cells (Comoli et al., 2006; Egli et al., 2007).
Immunosuppression regimens target lymphocytes, the major IFN-γ-producing cells, and
prevent the activation and proliferation of these cells. In Chapter II, we describe the
inhibitory effect of IFN-γ on BKV replication. A reduction in levels of this cytokine is
likely a key factor that allows BKV reactivation. In Chapter IV, we discuss future
research plans to examine the specific factors involved in this regulation. Transcription
factor binding site prediction programs were of limited help in identifying potential
responsive elements since IFN-γ-mediated inhibition of viral replication was strongest at
late stages of infection, when the effects of IFN-γ signaling are robust and widespread, as
a result of the activation and induction of many signaling components.
While IFN-γ is a well-known antiviral cytokine, it is not frequently shown to have
direct inhibitory activities on viral gene expression. It is far more common for the type I
interferon signaling pathway to mediate direct effects on viral replication, while IFN-γ
signaling typically regulates the cellular immune response and activates cytotoxic T
lymphocytes to kill virus-infected cells. In our studies, IFN-α did not significantly affect
BKV gene expression, suggesting that the virus is able to somehow avoid effectors of the
type I pathway. It is possible that transcription factors induced or activated by IFN-α/β
signaling may be unable to regulate the BKV promoter due to a lack of the appropriate
binding sites. Alternatively, expression of an early viral protein, likely TAg based on
similar findings for SV40 and Py (Swaminathan et al., 1996; Weihua et al., 1998), may
block the IFN-α/β signaling pathway. This explanation correlates with our observations
that BKV infection does not activate any innate immune response in kidney epithelial
145
cells (J. Abend, J. Low, and M. Imperiale, unpublished data), and may also partially
explain how BKV establishes persistence.
We hypothesize that a key factor mediating reactivation is intense
immunosuppresion of transplant patients. The immunosuppressive regimen during the
first year post-renal transplant is typically at higher doses than during later years posttransplant. Thus, it is possible that patients receive more immunosuppressant than is
necessary to maintain the balance between preventing graft rejection and controlling viral
reactivation. In addition, there may be a subset of transplant patients that naturally have
lower levels of cytokine production, and therefore would be more sensitive to
immunosuppression. Pravica et al. (1999) identified a polymorphism in the IFN-γ
promoter that results in higher levels of IFN-γ production. The polymorphism relates to
the number of consecutive CA repeats, and the allele (12 CA repeats, designated allele 2)
confers high production of IFN-γ in either homozygous or heterozygous individuals.
Allele 2 is more frequent than any other allele, with 75% of the population having at least
one copy (Pravica et al., 1999). Thus, 25% of the population is negative for allele 2 and
consequently has approximately two-fold lower levels of IFN-γ production, based on
stimulation of PBMCs in vitro. While this may not sound impressive, the effect in vivo
and in the presence of immunosuppressive drugs may be more dramatic. The frequency
of low IFN-γ producers seems to roughly correlate with the frequency of PVN and HC: if
25% of the transplant population is negative for allele 2, then it is reasonable to expect
about half of these patients (10%) have another risk factor, such as age gender or donor
seropositivity, and develop PVN.
146
In addition to polymorphisms that alter IFN-γ production, there are also
polymorphisms that determine production levels of other relevant cytokines, for example,
TGF-β (Grainger et al., 1999). The IL-12 gene has a single nucleotide polymorphism
that reduces the production of the p40 subunit of IL-12 (Stanilova and Miteva, 2005), and
has recently been reported as a risk factor for human cytomegalovirus infection after
kidney transplantation (Hoffmann et al., 2008). IL-18 also has promoter polymorphisms
that confer levels of expression (Giedraitis et al., 2001) and are implicated in disease
outcome. IL-12 and IL-18 (also known as the IFN-γ-inducing factor) are the major
activators of IFN-γ production, and can act alone or in synergy. Polymorphisms that
confer low levels of expression of these cytokines could be associated with a higher risk
of BKV reactivation. It would be interesting and useful to determine if there is a
correlation between these and other polymorphisms, and progression to HC or PVN.
Patients could easily be screened for such polymorphisms prior to transplantation to
determine if there is a high risk for reactivation. In this situation, the immunosuppressive
regimens could be adjusted and monitored.
Although immunosuppression is important, the infrequency of BKV-associated
disease in non-renal solid organ transplant recipients and patients with AIDS indicates
that there are other conditions required for reactivation. One such kidney-specific risk
factor could be damage to the graft as a result of ischemia and reperfusion during
transplantation. Ischemia/reperfusion events may activate factors within BKV-infected
cells that could promote viral reactivation, and since the kidney is the major site of BKV
persistence, these events would only affect renal transplant patients.
147
Conclusions
Based on the discussions in this chapter, a proposed model for the replication of
archetype virus (left) and the factors that determine the prevalence of rearranged virus in
the blood (right) is depicted in Figure 5.1. First, the archetype virus establishes a
persistent infection in kidney epithelial cells but is only able to express very low levels of
TAg that do not support replication. Random rearrangements produce a viral genome
that has a highly active early region promoter but is defective for late gene expression,
either by disruption of the coding region or an essential element of the late region
promoter. The resulting high levels of TAg then promote replication of archetype virus
without producing rearranged virus. The second part of this figure shows archetype virus
release by a cell-mediated egress pathway, which leaves the infected cell intact. The
slower kinetics of archetype replication in healthy individuals may promote the use of
this pathway, and egress may be directed to the apical side of the epithelial cell resulting
in preferential release in urine. The faster, more robust replication of rearranged strains
may prevent utilization of this pathway, and instead result in a viral lytic burst.
Destruction of the infected cell may promote spread to surrounding cells, in particular
vascular endothelium, resulting in a preference for the release of rearranged strains into
the bloodstream. It is likely that archetype strains also cause cell lysis under certain
conditions of robust replication, as these viruses are found both in blood and urine.
In conclusion, this dissertation has provided an overview of BKV biology and
pathogenesis, and a description of our findings on cytokine-mediated regulation of viral
gene expression and replication. We have also shown preliminary data on the mechanism
of IFN-γ-mediated regulation and the replication of archetype strains in culture, and
148
proposed further studies. Finally, we discussed the major questions in the field of BKV
reactivation in transplant patients. In the absence of effective antiviral therapies for
BKV, a better understanding of persistence and factors that drive reactivation will
provide insight into better options for the treatment and prevention of PVN and HC.
149
Figure 5.1. Proposed model for replication of archetype virus and factors that
determine the prevalence of rearranged virus in the blood. Replication of archetype
virus (left): archetype virus persistence is established in a cell (genomes shown by yellow
circles). Weak TAg expression (thin arrow) does not support replication, however,
random rearrangements produce a viral genome (purple circle) that has robust early
promoter activity but is unable to produce VP1 and thus unable to propagate. The high
levels of TAg (thick purple arrow) are utilized by the archetype virus to facilitate
replication. Virus release pathways (right): archetype virus utilizes a cell-mediated
egress pathway for progeny release, leaving the infected cell intact. Rearranged virus
(genomes shown by green circles), replicates robustly and overwhelms the cell with
progeny, leading to a viral lytic burst (green star shape). Arch, archetype; rearr,
rearranged.
150
References
Agostini, H. T., Ryschkewitsch, C. F., and Stoner, G. L. (1996). Genotype profile of
human polyomavirus JC excreted in urine of immunocompetent individuals. J
Clin Microbiol 34, 159-164.
Asselineau, D., and Pruniera, M. (1984). Reconstruction of 'simplified' skin: control of
fabrication. Br J Dermatol 27, 219-222.
Bendiksen, S., Rekvig, O. P., Van Ghelue, M., and Moens, U. (2000). VP1 DNA
sequences of JC and BK viruses detected in urine of systemic lupus
erythematosus patients reveal no differences from strains expressed in normal
individuals. J Gen Virol 81, 2625-2633.
Bettelli, E., Korn, T., Oukka, M., and Kuchroo, V. K. (2008). Induction and effector
functions of T(H)17 cells. Nature 453, 1051-1057.
Blank, C. A., Anderson, D. A., Beard, M., and Lemon, S. M. (2000). Infection of
polarized cultures of human intestinal epithelial cells with Hepatitis A virus:
vectorial release of progeny virions through apical cellular membranes. J Virol
74, 6476-6484.
Blau, D. M., and Compans, R. W. (1995). Entry and release of measles virus are
polarized in epithelial cells. Virology 210, 91-99.
Boldorini, R., Veggiani, C., Amoruso, E., Allegrini, S., Miglio, U., Paganotti, A.,
Ribaldone, R., and Monga, G. (2008). Latent human polyomavirus infection in
pregnancy: investigation of possible transplacental transmission. Pathology 40,
72-77.
Brown, K. A., Pietenpol, J. A., and Moses, H. L. (2007). A tale of two proteins:
differential roles and regulation of Smad2 and Smad3 in TGF-beta signaling. J
Cell Biochem 101, 9-33.
Chodosh, J., Gan, Y. J., Holder, V. P., and Sixbey, J. W. (2000). Patterned entry and
egress by Epstein-Barr virus in polarized CR2-positive epithelial cells. Virology
266, 387-396.
Clayson, E. T., Brando, L. V. J., and Compans, R. W. (1989). Release of simian virus 40
virions from epithelial cells is polarized and occurs without cell lysis. J Virol 63,
2278-2288.
Coleman, D. V., Wolfendale, M. R., Daniel, R. A., Dhanjal, N. K., Gardner, S. D.,
Gibson, P. E., and Field, A. M. (1980). A prospective study of human
polyomavirus infection in pregnancy. J Infect Dis 142, 1-8.
151
Comoli, P., Binggeli, S., Ginevri, F., and Hirsch, H. H. (2006). Polyomavirus-associated
nephropathy: update on BK virus-specific immunity. Transpl Infect Dis 8, 86-94.
Doerries, K., Vogel, E., Gunther, S., and Czub, S. (1994). Infection of human
polyomaviruses JC and BK in peripheral blood leukocytes from
immunocompetent individuals. Virology 198, 59-70.
Egli, A., Binggeli, S., Bodaghi, S., Dumoulin, A., Funk, G. A., Khanna, N., Leuenberger,
D., Gosert, R., and Hirsch, H. H. (2007). Cytomegalovirus and polyomavirus BK
posttransplant. Nephrol Dial Transplant 22, viii72-viii82.
Feng, X. H., and Derynck, R. (2005). Specificity and versatility in TGF-beta signaling
through Smads. Annu Rev Cell Dev Biol 21, 659-693.
Giannopoulou, M., Iszkula, S. C., Dai, C., Tan, X., Yang, J., Michalopoulos, G. K., and
Liu, Y. (2006). Distinctive roles for Stat3 and Erk-1/2 activation in mediating
interferon-gamma inhibition of TGF-beta1 action. Am J Physiol Renal Physiol
290, F1234-F1240.
Giedraitis, V., He, B., Huang, W. X., and Hillert, J. (2001). Cloning and mutation
analysis of IL-18 promoter: a possible role of polymorphisms in expression
regulation. J Neuroimmunol 112, 146-152.
Gorelik, L., and Flavell, R. A. (2002). Transforming growth factor-beta in T-cell biology.
Nat Rev Immunol 2, 46-53.
Gosert, R., Rinaldo, C. H., Funk, G. A., Egli, A., Ramos, E., Drachenberg, C. B., and
Hirsch, H. H. (2008). Polyomavirus BK with rearranged noncoding control region
emerge in vivo in renal transplant patients and increase viral replication and
cytopathology. J Exp Med 205, 841-852.
Grainger, D. J., Heathcote, K., Chiano, M., Snieder, H., Kemp, P. R., Metcalfe, J. C.,
Carter, N. D., and Spector, T. D. (1999). Genetic control of the circulating
concentration of transforming growth factor type beta1. Hum Mol Genet 8, 93-97.
Han Lee, E. D., Kemball, C. C., Wang, J., Dong, Y., Stapler, D. C., Hamby, K. M.,
Gangappa, S., Newell, K. A., Pearson, T. C., Lukacher, A. E., and Larsen, C. P.
(2006). A mouse model for polyomavirus-associated nephropathy of kidney
transplants. Am J Transplant 6, 913-922.
Hara, K., Sugimoto, C., Kitamura, T., Aoki, N., Taguchi, F., and Yogo, Y. (1998).
Archetype JC virus efficiently replicates in COS-7 cells, simian cells
constitutively expressing simian virus 40 T antigen. J Virol 72, 5335-5342.
152
Hirsch, H. H., Knowles, W., Dickenmann, M., Passweg, J., Klimkait, T., Mihatsch, M. J.,
and Steiger, J. (2002). Prospective study of polyomavirus type BK replication and
nephropathy in renal-transplant recipients. N Engl J Med 347, 488-496.
Hoffmann, T. W., Halimi, J. M., Buchler, M., Velge-Roussel, F., Goudeau, A., Najjar, A.
A., Boulanger, M. D., Houssaini, T. S., Marliere, J. F., Lebranchu, Y., and Baron,
C. (2008). Association between a polymorphism in the IL-12p40 gene and
cytomegalovirus reactivation after kidney transplantation. Transplantation 85,
1406-1411.
Howley, P. M., and Lowy, D. R. (2007). Papillomaviruses. Fifth ed. In "Fields Virology"
(D. M. Knipe, and P. M. Howley, Eds.), Vol. 2, 2 vols. Lippincott Williams &
Wilkins, Philadelphia, pp. 2299-2354.
Imperiale, M. J., and Major, E. O. (2007). Polyomaviruses. Fifth ed. In "Fields Virology"
(D. M. Knipe, and P. M. Howley, Eds.), Vol. 2, 2 vols. Lippincott Williams &
Wilkins, Philadelphia, pp. 2263-2298.
Jeong, B. H., Lee, K. H., Choi, E. K., Kim, K., and Kim, Y. S. (2004). Genotyping of the
JC virus in urine samples of healthy Korean individuals. J Med Virol 72, 281-289.
Kato, A., Kitamura, T., Takasaka, T., Tominaga, T., Ishikawa, A., Zheng, H. Y., and
Yogo, Y. (2004). Detection of the archetypal regulatory region of JC virus from
the tonsil tissue of patients with tonsillitis and tonsilar hypertrophy. J Neurovirol
10, 244-249.
Khalili, K., Gordon, J., and White, M. K. (2006). The polyomavirus, JCV, and its
involvement in human disease. Adv Exp Med Biol 577, 274-287.
Khanna, A., Cairns, V., and Hosenpud, J. D. (1999a). Tacrolimus induces increased
expression of transforming growth factor-beta1 in mammalian lymphoid as well
as nonlymphoid cells. Transplantation 67, 614-619.
Khanna, A. K., Cairns, V. R., Becker, C. G., and Hosenpud, J. D. (1999b). Transforming
growth factor (TGF)-beta mimics and anti-TGF-beta antibody abrogates the in
vivo effects of cyclosporine: demonstration of a direct role of TGF-beta in
immunosuppression and nephrotoxicity of cyclosporine. Transplantation 67, 882889.
Li, M. O., Wan, Y. Y., Sanjabi, S., Robertson, A. K., and Flavell, R. A. (2006).
Transforming growth factor-beta regulation of immune responses. Annu Rev
Immunol 24, 99-146.
Loeber, G., and Doerries, K. (1988). DNA rearrangements in organ-specific variants of
polyomavirus JC strain GS. J Virol 62, 1730-1735.
153
Low, J., Humes, H. D., Szczypka, M., and Imperiale, M. (2004). BKV and SV40
infection of human kidney tubular epithelial cells in vitro. Virology 323, 182-188.
Markowitz, R. B., Eaton, B. A., Kubik, M. F., Latorra, D., McGregor, J. A., and Dynan,
W. S. (1991). BK virus and JC virus shed during pregnancy have predominantly
archetypal regulatory regions. J Virol 65, 4515-4519.
Martin, J. D., King, D. M., Slauch, J. M., and Frisque, R. J. (1985). Differences in
regulatory sequences of naturally occurring JC virus variants. J Virol 53, 306-311.
McCance, D. J., Kopan, R., Fuchs, E., and Laimins, L. A. (1988). Human papillomavirus
type 16 alters human epithelial cell differentiation in vitro. Proc Natl Acad Sci U
S A 85, 7169-7173.
McMorrow, T., Gaffney, M. M., Slattery, C., Campbell, E., and Ryan, M. P. (2005).
Cyclosporine A induced epithelial-mesenchymal transition in human renal
proximal tubular epithelial cells. Nephrol Dial Transplant 20, 2215-2225.
Messam, C. A., Hou, J., Gronostajski, R. M., and Major, E. O. (2003). Lineage pathway
of human brain progenitor cells identified by JC virus susceptability. Ann Neurol
53, 636-646.
Moens, U., Subramaniam, N., Johansen, B., Johansen, T., and Traavik, T. (1994). A
steroid homone response unit in the late leader of the noncoding control region of
the human polyomavirus BK confers enhanced host cell permissivity. J Virol 68,
2398-2408.
Monaco, M. C., Sabath, B. F., Durham, L. C., and Major, E. O. (2001). JC virus
multiplication in human hematopoietic progenitor cells requires the NF-1 class D
transcription factor. J Virol 75, 9687-9695.
Monini, P., Rotola, A., Di Luca, D., De Lellis, L., Chiari, E., Corallini, A., and Cassai, E.
(1995). DNA rearrangements impairing BK virus productive infection in urinary
tract tumors. Virology 214, 273-279.
Negrini, M., Sabbioni, S., Arthur, R. R., Castagnoli, A., and Barbanti-Brodano, G.
(1991). Prevalence of the archetypal regulatory region and sequence
polymorphisms in nonpassaged BK virus variants. J Virol 65, 5092-5095.
Pravica, V., Asderakis, A., Perrey, C., Hajeer, A., Sinnott, P. J., and Hutchinson, I. V.
(1999). In vitro production of IFN-gamma correlates with CA repeat
polymorphism in the human IFN-gamma gene. Eur J Immunogenet 26, 1-3.
Priftakis, P., Bogdanovic, G., Kalantari, M., and Dalianis, T. (2001). Overrepresentation
of point mutations in the Sp1 site of the non-coding control region of BK virus in
154
bone marrow transplanted patients with haemorrhagic cystitis. J Clin Virol 21, 17.
Rahimi, R. A., and Leof, E. B. (2007). TGF-beta signaling: a tale of two responses. J Cell
Biochem 102, 593-608.
Rinaldo, C. H., Hansen, H., and Traavik, T. (2005). Human endothelial cells allow
passage of an archetypal BK virus (BKV) strain-a tool for cultivation and
functional studies of natural BKV strains. Arch Virol 150, 1449-1458.
Rinaldo, C. H., Traavik, T., and Hey, A. (1998). The agnogene of the human
polyomavirus BK is expressed. J Virol 72, 6233-6236.
Rubinstein, R., Schoonakker, B. C., and Harley, E. H. (1991). Recurring theme of
changes in the transcriptional control region of BK virus during adaptation to cell
culture. J Virol 65, 1600-1604.
Sharma, P. M., Gupta, G., Vats, A., Shapiro, R., and Randhawa, P. S. (2007).
Polyomavirus BK non-coding control region rearrangements in health and
disease. J Med Virol 79, 1199-1207.
Shihab, F. S., Andoh, T. F., Tanner, A. M., Noble, N. A., Border, W. A., Franceschini,
N., and Bennett, W. M. (1996). Role of transforming growth factor-beta 1 in
experimental chronic cyclosporine nephropathy. Kidney Int 49, 1141-1151.
Sock, E., Renner, K., Feist, D., Leger, H., and Wegner, M. (1996). Functional
comparison of PML-type and archetype strains of JC virus. J Virol 70, 15121520.
Stanilova, S., and Miteva, L. (2005). Taq-I polymorphism in 3'UTR of the IL-12B and
association with IL-12p40 production from human PBMC. Genes Immun 6, 364366.
Steinman, L. (2007). A brief history of T(H)17, the first major revision in the
T(H)1/T(H)2 hypothesis of T cell-mediated tissue damage. Nat Med 13, 139-145.
Sugimoto, C., Hara, K., Taguchi, F., and Yogo, Y. (1989). Growth efficiency of naturally
occurring BK virus variants in vivo and in vitro. J Virol 63, 3195-3199.
Sundsfjord, A., Flaegstad, T., Flo, R., Spein, A. R., Pedersen, M., Permin, H., Julsrud, J.,
and Traavik, T. (1994). BK and JC viruses in human immunodeficiency virus
type 1-infected persons: prevalence, excretion, viremia, and viral regulatory
regions. J Infect Dis 169, 485-490.
155
Sundsfjord, A., Johansen, T., Flaegstad, T., Moens, U., Villand, P., Subramani, S., and
Traavik, T. (1990). At least two types of control regions can be found among
naturally occurring BK virus strains. J Virol 64, 3864-3871.
Sundsfjord, A., Osei, A., Rosenqvist, H., Van Ghelue, M., Silsand, Y., Haga, H. J.,
Rekvig, O. P., and Moens, U. (1999). BK and JC viruses in patients with systemic
lupus erythematosus: prevalent and persistent BK viruria, sequence stability of the
viral regulatory regions, and nondetectable viremia. J Infect Dis 180, 1-9.
Swaminathan, S., Rajan, P., Savinova, O., Jagus, R., and Thimmapaya, B. (1996). Simian
virus 40 large-T bypasses the translational block imposed by the phosphorylation
of eIF-2alpha. Virology 219, 321-323.
Takasaka, T., Goya, N., Tokumoto, T., Tanabe, K., Toma, H., Ogawa, Y., Hokama, S.,
Momose, A., Funyu, T., Fujioka, T., Omori, S., Akiyama, H., Chen, Q., Zheng, H.
Y., Ohta, N., Kitamura, T., and Yogo, Y. (2004). Subtypes of BK virus prevalent
in Japan and variation in their transcriptional control region. J Gen Virol 85,
2821-2827.
Tato, C. M., and O'Shea, J. J. (2006). Immunology: what does it mean to be just 17?
Nature 441, 166-168.
ter Schegget, J., Sol, C. J. A., Wouters Baan, E., van der Noordaa, J., and van Ormondt,
H. (1985). Naturally occurring BK virus variants (JL and Dik) with deletions in
the putative early enhancer-promoter sequences. J Virol 53, 302-305.
Tominaga, T., Yogo, Y., Kitamura, T., and Aso, Y. (1992). Persistence of archetypal JC
virus DNA in normal renal tissue derived from tumor-bearing patients. Virology
186, 736-741.
Trofe, J., Roy-Chaudhury, P., Gordon, J., Mutema, G., Cavallo, T., Cardi, M., Austin, J.,
Goel, S., Rogers, C., Boardman, R., Clippard, M., Alloway, R., Alexander, J. W.,
Metze, T., Goodman, H., Hanaway, M., Munda, R., Buell, J. F., Peddi, R., Safdar,
S., Wadih, G., Huang, S., Fidler, J., Khalili, K., and Woodle, E. S. (2003). Study
3: early steriod cessation-avoidance regimens are associated with a lower
incidence of polyomavirus nephropathy compared with steriod-based
immunosuppression in kidney transplant recipients. Transplant Rev 17, S52-S53.
Ulloa, L., Doody, J., and Massague, J. (1999). Inhibition of transforming growth factorbeta/SMAD signalling by the interferon-gamma/STAT pathway. Nature 397, 710713.
Watanabe, S., and Yoshiike, K. (1982). Change of DNA near the origin of replication
enhances the transforming capacity of human papovavirus BK. J Virol 42, 978985.
156
Watanabe, S., and Yoshiike, K. (1986). Evolutionary changes of transcriptional control
region in a minute-plaque viable deletion mutant of BK virus. J Virol 59, 260266.
Weihua, X., Ramanujam, S., Lindner, D. J., Kudaravalli, R. D., Freund, R., and
Kalvakolanu, D. V. (1998). The polyoma virus T antigen interferes with
interferon-inducible gene expression. Proc Natl Acad Sci U S A 95, 1085-1090.
Yogo, Y., Kitamura, T., Sugimoto, C., Ueki, T., Aso, Y., Hara, K., and Taguchi, F.
(1990). Isolation of a possible archetypal JC virus DNA sequence from
nonimmunocompromised individuals. J Virol 64, 3139-3143.
Yogo, Y., and Sugimoto, C. (2001). The archetype concept and regulatory region
rearrangement. In "Human polyomaviruses: molecular and clinical perspectives"
(K. Khalili, and G. L. Stoner, Eds.). Wiley-Liss, Inc, New York, pp. 127-148.
157